SlideShare a Scribd company logo
1 of 24
Download to read offline
Themed Issue: New avenues in cancer prevention and treatment (BJP 75th Anniversary)
C O M M I S S I O N E D R E V I E W A R T I C L E - T H E M E D I S S U E
The importance of Ras in drug resistance in cancer
Fiona M. Healy1
| Ian A. Prior2
| David J. MacEwan1
1
Department of Pharmacology and
Therapeutics, Institute of Systems, Molecular
and Integrative Biology (ISMIB), University of
Liverpool, Liverpool, UK
2
Department of Molecular Physiology and Cell
Signalling, Institute of Systems, Molecular and
Integrative Biology (ISMIB), University of
Liverpool, Liverpool, UK
Correspondence
Prof. David J. MacEwan, Department of
Pharmacology and Therapeutics, Institute of
Systems, Molecular and Integrative Biology
(ISMIB), University of Liverpool, Liverpool L69
3GE, UK.
Email: macewan@liverpool.ac.uk
In this review, we analyse the impact of oncogenic Ras mutations in mediating cancer
drug resistance and the progress made in the abrogation of this resistance, through
pharmacological targeting. At a physiological level, Ras is implicated in many cellular
proliferation and survival pathways. However, mutations within this small GTPase
can be responsible for the initiation of cancer, therapeutic resistance and failure, and
ultimately disease relapse. Often termed “undruggable,” Ras is notoriously difficult to
target directly, due to its structure and intrinsic activity. Thus, Ras-mediated drug
resistance remains a considerable pharmacological problem. However, with advances
in both analytical techniques and novel drug classes, the therapeutic landscape
against Ras is changing. Allele-specific, direct Ras-targeting agents have reached
clinical trials for the first time, indicating there may, at last, be hope of targeting such
an elusive but significant protein for better more effective cancer therapy.
LINKED ARTICLES: This article is part of a themed issue on New avenues in cancer
prevention and treatment (BJP 75th Anniversary). To view the other articles in this
section visit http://onlinelibrary.wiley.com/doi/10.1111/bph.v179.12/issuetoc
K E Y W O R D S
acquired resistance, cancer stem cells, Drug resistance, intrinsic resistance, MAPK, PI3K
1 | INTRODUCTION
Resistance to conventional therapeutic agents is an increasingly con-
cerning issue across all areas of disease, including cancer. Whilst the
heterogeneous nature of cancer means there are many mechanisms
resulting in drug resistance, Ras mutations underpin resistance to a
variety of therapies (Hobbs et al., 2016; Prior et al., 2012, 2020).
Oncogenic mutations in this small GTPase, which occur in approxi-
mately 19% of all cancers, cause constitutive activation of prolifera-
tive and survival pathways (Prior et al., 2020). This can abrogate the
effects of standard chemotherapy and newer receptor-targeted
therapies. Examples of such resistance is seen across a wide range of
cancers, including metastatic colorectal cancer (mCRC), non-small cell
lung cancer (NSCLC), pancreatic cancer, acute myeloid leukaemia
(AML) and basal cell carcinoma (Li et al., 2004; McMahon et al., 2019;
Misale et al., 2012; Tao et al., 2014). Thus, there is a distinct clinical
need to target Ras pharmacologically. Whilst this has been particularly
challenging due to structural difficulties and very high levels of intrin-
sic activity, significant developments have been made within the last
decade. Allele-specific, direct Ras-targeting reaching clinical trials pre-
sent a new era of potential Ras therapeutics and increase the likeli-
hood of overcoming this resistance mechanism.
Abbreviations: EGFR, epidermal growth factor receptor; FLT3, fms related receptor tyrosine kinase 3; GAP, GTPase-activating protein; GEF, guanine nucleotide exchange factor; lncRNAs, long
non-coding RNAs; MEK, MAPK kinase; RTK, receptor tyrosine kinase; SOS, son-of-sevenless; SOS1, SOS Ras/Rac guanine nucleotide exchange factor 1; TKI, tyrosine kinase inhibitor.
Received: 17 December 2020 Revised: 10 February 2021 Accepted: 21 February 2021
DOI: 10.1111/bph.15420
This is an open access article under the terms of the Creative Commons Attribution License, which permits use, distribution and reproduction in any medium,
provided the original work is properly cited.
© 2021 The Authors. British Journal of Pharmacology published by John Wiley & Sons Ltd on behalf of British Pharmacological Society.
2844 Br J Pharmacol. 2022;179:2844–2867.
wileyonlinelibrary.com/journal/bph
2 | WHAT IS DRUG RESISTANCE?
There are two general classes of resistance, intrinsic and acquired.
Intrinsic resistance occurs due to overt pre-existing factors, including
variations in protein expression levels (such as increased expression of
the P-gp [MDR1] transporter), epigenetic modifications (including by
the long non-coding RNA (lncRNA) HAND2-AS1 and chromatin modi-
fier Jarid1A (lysine demethylase 5A)) and somatic mutations (such as
in Ras) (Burrell et al., 2013; Gruber et al., 2012; Marusyk et al., 2012;
Sharma et al., 2010). Changes conferring drug resistance often co-
exist, resulting in a resistance heterogeneity similar to that seen in the
original disease itself (Gerlinger et al., 2012; Ramirez et al., 2016).
Acquired resistance is thought to occur for a number of reasons,
including through pre-existing (but initially undetectable) and de novo
mutations (Bhaduri et al., 2020; Russo et al., 2019). This is often
identified weeks to months after treatment has commenced (Santoni-
Rugiu et al., 2019). In recent times, the concept of disease clonal
heterogeneity has provided greater insight into the causes of acquired
resistance, suggesting that chemoresistance and disease relapse occur
as a result of minor subclonal populations. Given that these likely
contribute to the heterogeneous nature of cancer, it seems likely that
these also contribute to disease re-emergence and relapse (Bonnet &
Dick, 1997; Gerlinger et al., 2012; Pattabiraman & Weinberg, 2014;
Roy & Cowden Dahl, 2018; Seth et al., 2019). Within these minor
clonal populations, mutations conferring drug resistance may exist at
the early stages of disease but remain dormant and undetectable upon
first presentation (Pietrantonio et al., 2017; Russo et al., 2018). When
the bulk of the cancer is eliminated as a result of initial chemotherapy
targeted at overt mutations, cells from this minor subclone proliferate
and become dominant in the tumour bulk (Jones et al., 2019;
McMahon et al., 2019).
A second, related, resistance mechanism involves a very rare
subpopulation of cells, known as cancer stem cells. Cancer stem cells
were first described in acute myeloid leukaemia but have since been
applied to many other cancers including (but not limited to) breast
cancer, colorectal cancer, myeloma and pancreatic adenocarcinoma
(Bonnet & Dick, 1997; Koury et al., 2017; Lapidot et al., 1994).
Cancer stem cells have unique properties compared to bulk tumour
cells. They are undifferentiated and have strong self-renewal and
proliferative capabilities and typically remain in a quiescent state,
thus avoiding chemotherapy targeted at rapidly dividing cells (as in
bulk tumour cells). However, they do have the proliferative capacity
to maintain and expand the tumour burden, as bulk tumour cells are
eliminated (Jordan et al., 2006). These “stemness characteristics”
render this subset of cells intrinsically resistant to chemotherapy,
with distinct immunophenotypic and molecular signatures (Bonnet &
Dick, 1997). This includes increased expression of efflux trans-
porter P-gp (MDR1) and the ability to repair damaged DNA, a
common method of inducing cancer cell death (Dean et al., 2005;
Pattabiraman & Weinberg, 2014).
Certain pathways up-regulated in cancer stem cells have been
linked to the increased self-renewal capacity seen in this subset of
cells. Key examples include the Wnt/β-catenin and Hedgehog
(Hh) pathways, as well as the Ras-dependent MAPK and PI3K/AKT
pathways. Of these, the Wnt/β-catenin pathway is perhaps the most
associated with cancer stem cells. Briefly, at a physiological level, Wnt
signalling is not active and β-catenin is ubiquitinated and sent for
proteasomal degradation following interaction with GSK3β, APC and
Axin. In cancer stem cells however, activation of Wnt signalling
inhibits formation of the β-catenin–GSK3β–APC–Axin complex,
stabilising β-catenin. This translocates to the nucleus whereby it
stimulates transcription of proliferative genes, including c-Myc. This
promotes proliferative signalling, increasing the self-renewal capacity
of cancer stem cells (Krausova & Korinek, 2014; Moon et al., 2014).
Although still ambiguous, it has been reported that β-catenin
stabilisation can promote Ras stabilisation and protects Ras against
proteasomal degradation (Jeong et al., 2018; Lee et al., 2018). This in
turn promotes MAPK and PI3K pathway signalling, further contribut-
ing to the self-renewal capacity of cancer stem cells.
Alternatively, up-regulation of the hedgehog (Hh) signalling
pathway has been shown in cancer stem cells. Whilst there are many
facets to this pathway, its activation can stimulate up-regulation of
stemness-associated transcription factors including NANOG, OCT4
and SOX2, thus promoting self-renewal (Boyer et al., 2005; Po
et al., 2010; Takahashi & Yamanaka, 2006; Zhu et al., 2019). Increased
activation and stabilisation of these genes (through either Hh signal-
ling or biochemical stresses) have been implicated in the dedifferentia-
tion of bulk tumour cells to cancer stem cells (Herreros-Villanueva
et al., 2013; Kumar et al., 2012).
Interest in epigenetic remodelling has grown considerably in the
last decade and it has considerable effects in promoting drug resis-
tance. lncRNAs are also implicated in survival of cancer stem cells. In
hepatocellular carcinoma, the lncRNA HAND2-AS1 is up-regulated
and stimulates the self-renewal capacity of cancer stem cells, through
activation of the bone morphogenetic protein (BMP) signalling path-
way. BMP activation has in turn been shown to induce
chemoresistance in other cancer models, such as lung cancer (Gruber
et al., 2012; Wang et al., 2015). Indeed, this pathway is also regulated
by the fusion gene CBFA2T3-GLIS2 and the Hh and JAK-STAT signal-
ling pathways (Gruber et al., 2012; Okada et al., 2014).
Taken together, it is evident that many signalling pathways and
associated factors play a critical role in mediating drug resistance.
Identifying commonalities between these pathways could present an
opportunity to reduce the incidence and impact of drug resistance.
Although the incidence of Ras mutations in cancer stem cells is rela-
tively low, Ras-mediated pathways have been implicated (Corces-
Zimmerman et al., 2014). In addition to the aforementioned β-catenin-
mediated stabilisation of Ras, stabilised OCT4 and SOX2 expression
has also been shown in response to AKT activation, by extracellular
biochemical and radiation stresses (Maiuthed et al., 2018; Park
et al., 2021). Indeed, activation of the MAPK pathway through the
Ras–Raf interaction promotes increased expression of SOX2 and
NANOG (Chan et al., 2018; Du et al., 2019). Furthermore, correlation
has been seen between up-regulation of the MAPK, PI3K and BMP
pathways in drug-resistant lung cancer (Wang et al., 2015). Thus, not
only can these pathways contribute to an increased cellular
HEALY ET AL. 2845
14765381,
2022,
12,
Downloaded
from
https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420,
Wiley
Online
Library
on
[11/06/2023].
See
the
Terms
and
Conditions
(https://onlinelibrary.wiley.com/terms-and-conditions)
on
Wiley
Online
Library
for
rules
of
use;
OA
articles
are
governed
by
the
applicable
Creative
Commons
License
proliferation rate in bulk cancer, but they can also assist with the self-
renewal and stemness properties evident in cancer stem cells. Perhaps
unsurprisingly, inhibition of the Hh signalling pathway, when com-
bined with inhibition of mTOR (downstream of AKT), has been seen
to eliminate cancer stem cells in pancreatic cancer (Mueller
et al., 2009; Prieur et al., 2017). Furthermore, MAPK kinase (MEK)
inhibition has been implicated in reduced pancreatic cancer stem cell
survival (Walter et al., 2019).
3 | KEY EXAMPLES OF DRUG RESISTANCE
IN CANCER
In recent years, resistance has been documented amongst many can-
cer therapeutics, including traditional chemotherapeutic agents, more
novel small molecule inhibitors and monoclonal antibodies (Caiola
et al., 2015; McMahon et al., 2019; Pietrantonio et al., 2017). The
origin of this resistance (intrinsic or acquired) varies considerably
between drugs and so advances have been made in treatment stratifi-
cation, based on a patient's mutational status, for some therapies. For
example, vemurafenib, the BRAF inhibitor, is restricted to melanoma
patients with the BRAF V600E mutation (Hopkins et al., 2019). Like-
wise, the monoclonal antibody panitumumab is only recommended
for patients with wild-type Ras (Amado et al., 2008). However, as
mentioned previously, some mutations are only detectable when
patients relapse, such as the emergence of fms-like TK 3 (FLT3) or
NRAS mutations in acute myeloid leukaemia patients (Man
et al., 2012; McMahon et al., 2019; Piloto et al., 2007; Smith
et al., 2017). Many of these mutations occur downstream of the site
targeted by the drug. For example, since Ras operates downstream of
the EGF receptor (EGFR), Ras mutations play a role in rendering
cetuximab and panitumumab (which bind and inhibit EGFR) ineffec-
tive as the constitutive activation is not inhibited by the drug
(De Roock et al., 2010; Li et al., 2015; Zhao et al., 2017). However,
some of these putative resistance-causing mutations can only be
detected at relapse, once they have expanded from only existing in a
minor, undetectable subclone (as they did at diagnosis). Thus, it is
difficult to predict which patients will develop these resistance
mutations, meaning initial treatment stratification is difficult. Thus,
combination therapy between these established agents and novel
agents targeting these common mutational sites may be a way
forward in preventing resistance before it occurs, an example of which
is the combination of BRAF and MEK inhibitors, vemurafenib and
cobimetinib (Hopkins et al., 2019).
4 | INTRODUCTION TO Ras
As eluded, Ras mutations underpin resistance to a variety of therapies.
KRAS is most commonly mutated of the three Ras isoforms and is
particularly frequently mutated in lung and pancreatic cancer (Moore,
Rosenberg, et al., 2020; Prior et al., 2020). However, patterns of Ras
mutation differ between cancers and NRAS is the most frequently
mutated RAS gene in acute myeloid leukaemia. Also, HRAS accounts
for most Ras mutations in head and neck squamous cell carcinoma
(Prior et al., 2020). Indeed, this variation carries varying prognoses of
Ras-mutated cancer, with KRAS mutations conferring the lowest over-
all survival 5 years post diagnosis (43%) and HRAS mutations confer-
ring the highest (63%) (Figure 1) (Cerami et al., 2012; Gao
et al., 2013).
Most oncogenic mutations in Ras occur in the residues G12, G13
and Q61, which are located in a region conserved between HRAS,
KRAS and NRAS (Figure 2a). They occur within the effector lobe,
which comprises the first 85 amino acids and is a region of complete
sequence identity between the different RAS isoforms. Most key
interaction sites occur within the effector lobe, including nucleotide
and effector interaction sites, as well as the switch regions that medi-
ate effector interactions (Figure 2b). There is 90% similarity in the
next 80 amino acids (allosteric lobe). The only considerable sequence
differences occur at the C terminal end of the protein, known as the
hypervariable region. This is the region in which post-translational
modifications occur and membrane-targeting sequences are found
(Hobbs et al., 2016). Initial attempts at direct pharmacological
targeting of Ras centred around inhibiting these post-translational
modifications, including farnesylation (Cox & Der, 2002; Whyte
et al., 1997). However, more recent attempts have considered the
effector lobe and structures within this to be better therapeutic
targets, as discussed later (Canon et al., 2019; Janes et al., 2018;
Ostrem et al., 2013).
Ras is activated or inactivated when bound to GTP or GDP,
respectively. Oncogenic mutations increase the frequency of GTP-
bound (active) Ras, through two key mechanisms. This is either
through decreasing the affinity of Ras for GTPase-activating proteins
(GAPs), which stimulate the intrinsic GTPase activity of Ras and there-
fore facilitate hydrolysis of GTP to GDP (off/inactivating mechanism)
or by decreasing the need for guanine nucleotide exchange factors
(GEFs), which mediate the on/activating mechanism (Smith
et al., 2013; Wittinghofer & Waldmann, 2000). These differences
depend on multiple factors including the specific amino acid mutation
and resulting protein conformation (Hunter et al., 2015; Miller &
Miller, 2012; Poulin et al., 2019; Prior et al., 2012; Smith et al., 2013).
Some mutations (including G12C) promote rapid cycling between
the inactive and active states, an area which has been exploited by a
range of novel Ras-targeting drugs (Fell et al., 2020; Hunter
et al., 2015; Patricelli et al., 2016).
5 | Ras SIGNALLING—REGULAR AND
ONCOGENIC
Ras is involved in key pathways regulating cell proliferation,
differentiation and sensitivity to apoptosis, perhaps the most notable
being the MAPK pathway and the PI3K/AKT pathway. KRAS has also
been implicated (albeit in a more indirect way) in a range of other
pathways, including the Wnt/β-catenin pathway and the nuclear
factor erythroid 2-related factor 2 (NRF2) pathway (DeNicola
2846 HEALY ET AL.
14765381,
2022,
12,
Downloaded
from
https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420,
Wiley
Online
Library
on
[11/06/2023].
See
the
Terms
and
Conditions
(https://onlinelibrary.wiley.com/terms-and-conditions)
on
Wiley
Online
Library
for
rules
of
use;
OA
articles
are
governed
by
the
applicable
Creative
Commons
License
et al., 2011; Ferino et al., 2020; Moon et al., 2014; Park et al., 2019;
Tao et al., 2014). These pathways are detailed briefly below and are
summarised in Figure 3.
Ligand binding to receptor tyrosine kininases (RTKs), such as
EGFR or FLT3, causes autophosphorylation of key tyrosine residues
on the RTK, after which adaptor proteins such as GRB2 bind to these
phosphorylated tyrosines via SH2 domains. Son-of-sevenless (SOS), a
Ras-guanine nucleotide exchange factor, then associates with GRB2
via two SH3 domains and ultimately facilitates the exchange of GDP
for GTP on RAS, thereby activating it (Freeman, 2000).
Following this, Ras stimulates recruitment of its serine–threonine
kinase effector, RAF, to the membrane, which binds through its Ras
binding domain and then proceeds to phosphorylate downstream
kinases including MEK and ERK (Marais et al., 1995; Molina &
Adjei, 2006). ERK then translocates from the cytoplasm to the
nucleus, causing the activation of many transcription factors, which
FIGURE 1 The impact of Ras alterations on disease. Data collected from 32 curated, non-redundant studies, comprising 10,967 patient
samples, as collated by the TCGA Pan-Cancer Atlas. (a) Ten-year profession-free survival (PFS) analysis for patients with and without Ras
alterations. Ras wild-type (blue) versus altered (red) overall survival. Ras wild-type median PFS = 65.88 months, Ras-altered median
PFS = 44.19 months. P < .01. (b) Ten-year PFS stratified by altered Ras isoform. KRAS (orange) median PFS = 36.72 months, NRAS (green)
median PFS = 45.67 months, HRAS (purple) median PFS = 74.89 months. P < .01. (c) Cancer types where KRAS is altered in at least 10% of cases.
(d) Cancer types where NRAS is altered in at least 10% of cases. (e) Cancer types where HRAS is altered in at least 10% of cases. Data obtained
from cBioPortal, all TCGA Pan-Cancer Atlas Studies
HEALY ET AL. 2847
14765381,
2022,
12,
Downloaded
from
https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420,
Wiley
Online
Library
on
[11/06/2023].
See
the
Terms
and
Conditions
(https://onlinelibrary.wiley.com/terms-and-conditions)
on
Wiley
Online
Library
for
rules
of
use;
OA
articles
are
governed
by
the
applicable
Creative
Commons
License
T
A
B
L
E
1
Summary
of
current
Ras
pathway
targeting
drugs,
at
varying
stages
of
development
Name
Drug
type
Target
Development
stage
Mode
of
inhibition
Selected
common
adverse
effects
References
Gilteritinib
Pyrazinecarboxamide,
small
molecule
inhibitor
FLT3-ITD,
FLT3-TKD,
AXL
FDA-approved
Binds
active
FLT3
(either
ITD
or
TKD),
inhibiting
constitutive
signalling
Acute
kidney
injury,
hypotension,
diarrhoea,
dizziness
Lee
et
al.
(2017);
Perl
et
al.
(2019)
Midostaurin
Indolocarbazole,
small
molecule
inhibitor
FLT3
FDA-approved
Binds
active
FLT3,
inhibiting
constitutive
signalling
Nausea,
neutropenia,
thrombocytopenia,
diarrhoea,
vomiting
Levis
(2017);
Stone
et
al.
(2012)
Crenolanib
Benzamidazole-derivative
small
molecule
inhibitor
FLT3-ITD,
FLT3-TKD
Phase
II
(NCT02400255)
Binds
active
FLT3,
inhibiting
constitutive
signalling
Nausea,
vomiting,
diarrhoea,
stomach
pain
Galanis
et
al.
(2014);
Marensi
et
al.
(2021);
Wu
et
al.
(2018)
Cetuximab
Chimeric
IgG1
chimeric
monoclonal
antibody
EGFR
FDA-approved
Competitively
binds
extracellular
domain
of
EGFR,
preventing
ligand-mediated
signalling
Oedema,
fatigue,
anorexia,
rash,
vomiting
Jonker
et
al.
(2007)
Panitumumab
Human
IgG2
monoclonal
antibody
EGFR
FDA-approved
Competitively
binds
extracellular
domain
of
EGFR,
inhibiting
EGFR
dimerization
and
autophosphorylation
Rash,
diarrhoea,
hypomagnesaemia
Van
Cutsem
et
al.
(2007)
BAY293
Quinazoline-based
small
molecule
inhibitor
SOS1
Preclinical
Disruption
of
the
KRAS–SOS1
interaction,
reducing
exchange
of
GDP
for
GTP
N/A
Hillig
et
al.
(2019)
SAH-SOS1
Stapled
peptide
fragment
SOS1
Preclinical
Binds
KRAS
(active
or
inactive)
in
the
SOS1
binding
pocket,
reducing
GTP
binding
to
KRAS
N/A
Leshchiner
et
al.
(2015)
BI1701963
Small
molecule
inhibitor
SOS1
Phase
I
(NCT04111458)
Binds
inside
the
catalytic
site
of
SOS1,
preventing
interaction
with
(and
activation
of)
KRAS
N/A
Gerlach
et
al.
(2020)
KYA1797K
Thiazolidine-based
small
molecule
inhibitor
Axin
Preclinical
Binds
to
the
RGS
domain
of
Axin,
causing
formation
of
the
β-catenin
destruction
complex,
subsequently
destabilising
Ras
N/A
Cha
et
al.
(2016);
Lee
et
al.
(2018)
Tipifarnib
Quinolinone
small
molecule
inhibitor
Farnesyltransferase
FDA-approved
Prevents
Ras
farnesylation
and
subsequent
trafficking
to
the
membrane
Thrombocytopenia,
anorexia,
anaemia,
nausea,
neutropenia
Duffy
and
Crown
(2021);
Lee
et
al.
(2020)
ARS-853
Acrylamide-based,
small
molecule
inhibitor
KRAS
G12C
Preclinical
Irreversible
binding
to
KRAS
G12C
SII-P
pocket,
inhibiting
exchange
of
GDP
for
GTP
N/A
Lito
et
al.
(2016);
Patricelli
et
al.
(2016)
2848 HEALY ET AL.
14765381,
2022,
12,
Downloaded
from
https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420,
Wiley
Online
Library
on
[11/06/2023].
See
the
Terms
and
Conditions
(https://onlinelibrary.wiley.com/terms-and-conditions)
on
Wiley
Online
Library
for
rules
of
use;
OA
articles
are
governed
by
the
applicable
Creative
Commons
License
T
A
B
L
E
1
(Continued)
Name
Drug
type
Target
Development
stage
Mode
of
inhibition
Selected
common
adverse
effects
References
ARS-1620
Acrylamide-based,
small
molecule
inhibitor
(S-atropisomer)
KRAS
G12C
Preclinical
Irreversible
binding
to
KRAS
G12C
SII-P
pocket,
inhibiting
exchange
of
GDP
for
GTP
N/A
Janes
et
al.
(2018)
AMG-510
Acrylamide-based,
small
molecule
inhibitor
KRAS
G12C
Phase
I/II
(NCT03600883)
Irreversible
binding
to
KRAS
G12C
SII-P
pocket,
inhibiting
exchange
of
GDP
for
GTP
Anaemia,
diarrhoea
Canon
et
al.
(2019);
Govindan
(2019)
MRTX849
Acrylamide-based,
small
molecule
inhibitor
KRAS
G12C
Phase
I
(NCT03785249)
Covalent
binding
in
SII-P
pocket,
inhibiting
exchange
of
GDP
for
GTP
Nausea,
diarrhoea,
fatigue,
hyponatremia
Fell
et
al.
(2020);
Hallin
et
al.
(2020a)
inRas37
Human
IgG1
internalising
and
PPI-interfering
monoclonal
antibody
pan-Ras
Preclinical
Competitive
inhibition
of
RAS–
effector
interaction
N/A
Shin
et
al.
(2020)
siG12D-
LODER
Long-acting
siRNA
KRAS
G12D
Phase
II
(NCT01676259)
siRNA-mediated
KRAS
G12D
silencing
Diarrhoea,
abdominal
pain,
nausea,
fatigue
Golan
et
al.
(2015);
Zorde
Khvalevsky
et
al.
(2013)
Dabrafenib
Sulphonamide-based
small
molecule
inhibitor
BRAF
(wild-type
and
V600-mutated)
FDA-approved
ATP-competitive
inhibitor,
binds
active
BRAF
and
thus
inhibits
downstream
effector
activation
Rash,
fever,
fatigue,
headache,
hypertension,
arthralgia
Bowyer
et
al.
(2015);
King
et
al.
(2013);
Rheault
et
al.
(2013)
Vemurafenib
Azaindole-derived
small
molecule
inhibitor
BRAF
V600E
FDA-approved
ATP-competitive
inhibitor,
binds
active
BRAF
and
thus
inhibits
downstream
receptor
activation
Skin
lesions
(including
squamous
cell
carcinoma),
arthralgia,
fatigue
Sharma
et
al.
(2012);
Tsai
et
al.
(2008)
Sorafenib
Biaryl-urea-based
small
molecule
inhibitor
RAF,
PDGFR,
VEGFR
amongst
others
FDA-approved
Binds
inactive
conformation
of
BRAF,
reducing
activation.
Also
sequesters
Raf
into
inactive
complexes
Diarrhoea,
weight
loss,
skin
reactions,
alopecia,
voice
changes
Adnane
et
al.
(2006);
Llovet
et
al.
(2008);
Marensi
et
al.
(2021);
Wan
et
al.
(2004);
Wilhelm
et
al.
(2006)
Cobimetinib
Carboxamide-based
small
molecule
inhibitor
RAF,
MEK
FDA-approved
Non-ATP-competitive
inhibitor,
binds
active
MEK,
inhibiting
downstream
ERK
activation
Diarrhoea,
rash,
fatigue,
arthralgia,
photosensitivity
Garnock-Jones
(2015);
Hatzivassiliou
et
al.
(2013);
Rice
et
al.
(2012)
Trametinib
Pyridopyrimidine
small
molecule
inhibitor
MEK
FDA-approved
ATP
non-competitive
kinase
inhibitor,
binds
the
kinase
suppressor
of
RAS–RAS
interface,
reducing
MEK
phosphorylation
Diarrhoea,
rash,
blurred
vision
Borthakur
et
al.
(2016);
Khan
et
al.
(2020)
LY3214996
Thiazolone-based
small
molecule
inhibitor
ERK
Phase
I
(NCT02857270)
ATP-competitive
inhibitor,
reversibly
binds
ERK1
and
ERK2,
causes
cell
cycle
arrest
in
G
1
and
initiates
apoptosis
Nausea,
vomiting,
diarrhoea,
fatigue,
blurred
vision
Bhagwat
et
al.
(2020);
Pant
et
al.
(2019)
(Continues)
HEALY ET AL. 2849
14765381,
2022,
12,
Downloaded
from
https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420,
Wiley
Online
Library
on
[11/06/2023].
See
the
Terms
and
Conditions
(https://onlinelibrary.wiley.com/terms-and-conditions)
on
Wiley
Online
Library
for
rules
of
use;
OA
articles
are
governed
by
the
applicable
Creative
Commons
License
go on to regulate a host of processes, including proliferation and dif-
ferentiation (Guo et al., 2020; Zhang & Liu, 2002).
Alternatively, activated Ras can promote the PI3K–AKT pathway.
This pathway also plays a key role in controlling survival, division and
metabolism and can be activated in many different ways (Castellano &
Downward, 2011; Mendoza et al., 2011). Ras binds and activates the
p110α catalytic subunit of PI3K, which in turn promotes transforma-
tion of the plasma membrane-bound lipid PIP2 to PIP3 and activation
of AKT by binding through its pleckstrin homology (PH) domain (Can-
tley, 2002; Castellano & Downward, 2011). AKT interacts with many
downstream effectors, including MDM2, BAX BAD and NF-κB. These
interactions regulate apoptosis (Cantley, 2002; Castellano & Down-
ward, 2011; Chang et al., 2003; Duronio, 2008). Furthermore, activa-
tion of FOXO by AKT promotes cellular metabolism (Engelman
et al., 2006). Given that the MAPK and PI3K pathways are both
strongly involved in controlling cell survival and death, it is clear that
aberrant signalling within these pathways will lead to cancer, of which
key hallmarks include sustaining proliferative signalling, enabling repli-
cative immortality and resisting cell death (Hanahan &
Weinberg, 2011).
Aside from these two classic Ras-dependent pathways, increasing
evidence suggests the implication of normal and mutant Ras function
in alternative mechanisms, such as redox homeostasis and stem cell
survival (Mukhopadhyay et al., 2020; Wu et al., 2019). The role of Ras
in cancer stem cells is yet to be fully elucidated. However, its crosstalk
with the Wnt/β-catenin pathway in tumorigenesis may suggest its
involvement in the maintenance of cancer stem cells (Moon
et al., 2014). It is understood that, in colorectal cancer, there is over-
activation of both the Wnt/β-catenin and Ras pathways, through
mutations within APC (a tumour suppressor) and KRAS, respectively
(Jeong et al., 2018). There is a synergistic effect seen with these
mutations (D'Abaco et al., 1996; Janssen et al., 2006; Jeong
et al., 2018; Margetis et al., 2017). Mutations within APC cause loss of
function of the tumour suppressor gene, whilst KRAS mutations lead
to phosphorylation of key tyrosine residues within β-catenin, causing
its accumulation within the cytoplasm. This ultimately increases acti-
vation of downstream Wnt pathway target genes, including REG4, a
marker of cancer stem cells (Hwang et al., 2020; Janssen et al., 2006).
Up-regulation of this pathway is highly associated with the increased
survival and plasticity properties seen in cancer stem cells and has
been seen in many cancers (Al-Hajj et al., 2003; Koury et al., 2017;
Lapidot et al., 1994). Taken together, it seems plausible that these
KRAS mutations may play a role in the protection of the minor subset
of cells, which likely have a key role in relapse.
Furthermore, Ras mutations can also be implicated in ROS
generation/detoxification. ROS are considered a “double-edged
sword,” capable of both helping and hindering cancer cells (Hayes &
McMahon, 2006; Wu et al., 2019). At physiological levels of ROS,
NRF2 binds Keap1 and is ubiquitinated and degraded on a regular
basis. However, upon detection of high levels of ROS (which is carci-
nogenic), NRF2 is unable to bind to Keap1 (due to conformational
changes in Keap1) and so translocates to the nucleus, where it acts as
a transcription factor for various downstream detoxification genes.
T
A
B
L
E
1
(Continued)
Name
Drug
type
Target
Development
stage
Mode
of
inhibition
Selected
common
adverse
effects
References
RBC8
Carbonitrile-based
small
molecule
inhibitor
RAL
Preclinical
Non-ATP-competitive
inhibitor,
binds
GDP-loaded
RAL,
preventing
activation
N/A
Walsh
et
al.
(2019);
Yan
et
al.
(2014)
Alpelisib
Aminothiazole-based
small
molecule
inhibitor
PI3Kα
FDA-approved
ATP-competitive
inhibitor,
binds
selectively
to
PI3Kα
Hyperglycaemia,
rash,
diarrhoea,
nausea,
decreased
appetite
André
et
al.
(2019);
Furet
et
al.
(2013)
Uprosertib
Thiophenecarboxemide-based
small
molecule
inhibitor
AKT
Phase
II
(NCT01902173)
ATP-competitive
inhibitor,
binds
AKT
and
reduces
downstream
signalling
Nausea,
vomiting,
diarrhoea,
rash
Gungor
et
al.
(2015);
Nitulescu
et
al.
(2016)
Everolimus
Macrocyclic
lactone-based
small
molecule
inhibitor
mTOR
FDA-approved
Complexes
with
FKBP12,
which
binds
mTOR
and
inhibits
activation
Leukopenia,
hypercholesterolaemia,
hyperlipidaemia
Dunn
and
Croom
(2006)
Abbreviations:
PPI,
protein–protein
interaction;
SOS,
son-of-sevenless;
2850 HEALY ET AL.
14765381,
2022,
12,
Downloaded
from
https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420,
Wiley
Online
Library
on
[11/06/2023].
See
the
Terms
and
Conditions
(https://onlinelibrary.wiley.com/terms-and-conditions)
on
Wiley
Online
Library
for
rules
of
use;
OA
articles
are
governed
by
the
applicable
Creative
Commons
License
FIGURE 2 2D and 3D representation of the structure of RAS. (a) Physiological binding domains. (b) Key structural domains (switch regions
and lobes), with mutational hotspots G12, G13 and Q61 indicated. Redrawn from Prior et al. (2012). 3D structures based on PDB 4DST
FIGURE 3 RAS-mediated pathways and associated inhibitors. Targets of small molecule inhibitors and monoclonal antibodies used across a
range of cancers to inhibit proliferative signalling and survival of cancer cells. Figure includes examples of compounds identified in vitro, those
which have progressed into trials and those which are approved. Further detail is provided in Table 1
HEALY ET AL. 2851
14765381,
2022,
12,
Downloaded
from
https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420,
Wiley
Online
Library
on
[11/06/2023].
See
the
Terms
and
Conditions
(https://onlinelibrary.wiley.com/terms-and-conditions)
on
Wiley
Online
Library
for
rules
of
use;
OA
articles
are
governed
by
the
applicable
Creative
Commons
License
Hence, the DNA of the healthy cells remains undamaged by ROS
(Basak et al., 2017; Gorrini et al., 2013). However, in cancer, the NRF2
pathway can be up-regulated to constitutively degrade ROS (induced
by chemotherapeutics), thereby conferring a protective effect to the
cancer cells, resulting in chemoresistance (Basak et al., 2017; Gorrini
et al., 2013). Therefore, the balance of NRF2 activation level is crucial.
In the last decade, it has been shown that KRAS G12D mutations can
increase NRF2 transcription, through activation of the TRE (12-O-
Tetradecanoylphorbol-13-acetate (TPA) response element) via the
MAPK pathway (DeNicola et al., 2011; Mukhopadhyay et al., 2020;
Shirazi et al., 2020; Tao et al., 2014). Thus, these KRAS mutations can
render the cancer cell more capable of coping with chemotherapy-
induced ROS, thereby mediating drug resistance.
6 | THE INVOLVEMENT OF Ras IN
CHEMOTHERAPY RESISTANCE
Should mutations occur as described above, it follows that one or
many of these pathways can be perturbed, leading to increased cell
proliferation, decreased cell death and promotion of cancer stem cells,
amongst other effects. The following scenarios illustrate some of
these key resistance mechanisms, highlighting the necessity for better
Ras targeting.
Platinum-based agents, such as cisplatin, carboplatin and
oxaliplatin, are used in the treatment of a variety of cancers, including
head and neck squamous cell carcinoma, testicular cancer and
non-small cell lung cancer (de Vries et al., 2020; Silva et al., 2019;
Weykamp et al., 2020). They are DNA intercalating agents that
interfere with RNA transcription and DNA replication, through cross-
linking of DNA. This results in the formation of DNA adducts, which
in turn drive the tumour cell to apoptosis. Cisplatin also induces
mitochondrial ROS, which further increase DNA damage and thus
increase the cytotoxic properties of the drug (Marullo et al., 2013;
Srinivas et al., 2019). However, there are many resistance mechanisms
associated with cisplatin, including the involvement of oncogenic
KRAS mutations (Caiola et al., 2015; DeNicola et al., 2011; Feldman
et al., 2014; Garassino et al., 2011). KRAS mutations were shown to
induce NRF2 pathway up-regulation in non-small cell lung cancer,
thereby decreasing cisplatin-induced ROS within the tumour cell, and
ultimately leading to decreased cell death (DeNicola et al., 2011). This
was supported by further work indicating oncogenic KRAS can induce
NRF2 gene transcription via the TPE response element, resulting in
the overactivation of the anti-oxidative stress pathway, rendering the
tumour cells resistant to cisplatin-induced ROS (Tao et al., 2014).
Furthermore, KRAS mutations can lead to hyperactivation of the
PI3K–AKT pathway, which is starting to be implicated as a cisplatin
resistance mechanism. As mentioned, up-regulation of the PI3K–
AKT–mTOR pathway can have multiple effects, including inhibition of
apoptosis and increased cell proliferation (de Vries et al., 2020).
Whilst there are other reasons for cisplatin resistance, Ras pathway
mutations are heavily implicated in the key mechanisms. Thus,
pharmacologically targeting Ras would provide an opportunity to
overcome many causes of this resistance.
Up-regulation of Ras-mediated pathways as a means of
chemoresistance is by no means restricted to cisplatin resistance
and is a common mechanism of resistance to TK inhibitors (TKIs),
such as those targeting RTKs including FLT3 and EGFR (Eberlein
et al., 2015; Massarelli et al., 2007; McMahon et al., 2019;
Ortiz-Cuaran et al., 2016; Piloto et al., 2007; Van Emburgh
et al., 2016). TKIs are used in a variety of cancers, including renal cell
carcinoma (RCC), colorectal cancer, acute myeloid leukaemia and non-
small cell lung cancer, to name a few examples. The mechanism of
action of TKIs involves inhibition of phosphorylation sites within the
protein, thereby preventing it exerting kinase activity on downstream
effectors (Ciardiello & Tortora, 2008; Yamaoka et al., 2018). However,
resistance to these can occur through two predominant mechanisms,
mutations within the RTK or mutations within downstream pathways
(McMahon et al., 2019; Piloto et al., 2007; Van Emburgh et al., 2016;
Yu et al., 2013). Given that Ras occurs downstream of these recep-
tors, any mutations within Ras will render the cell resistant to the TKI.
For example, studies have shown KRAS mutations render patients
resistant to gefitinib, used to treat non-small cell lung cancer (Pao
et al., 2005; Zhao et al., 2017). In a similar way, the treatment of
colorectal cancer with anti-EGFR monoclonal antibodies cetuximab or
panitumumab is only successful in a subset of patients, with many
eventually developing resistance (Pietrantonio et al., 2017). This has
been attributed to Ras mutations and variations in the EGFR extracel-
lular domain, which reduce antibody binding efficiency, ultimately
initiating relapse (Van Emburgh et al., 2016). Although cetuximab and
panitumumab are only prescribed to Ras wild-type patients, emer-
gence of mutations from undetectable, pre-existing clones can give
rise to resistance in this way, as evidenced through analysis of circu-
lating tumour DNA (ctDNA) (Amirouchene-Angelozzi et al., 2017; Diaz
et al., 2012; Misale et al., 2012). The order in which these mutations
develop/emerge is likely important in understanding (and ultimately
targeting) the process of relapse. Ras mutations often develop earlier
than EGFR variations and typically confer poorer prognosis (Van
Emburgh et al., 2016). Therefore, combatting these Ras mutations
would not only improve prognosis of Ras-mutated patients but also
provide a second therapeutic option for those that go on to develop
extracellular domain variations.
Resistance to FLT3-TKIs is also a highly prevalent issue. It is well
documented that 20–30% of acute myeloid leukaemia patients have
an internal tandem duplication in the FLT3 RTK (FLT3-ITD) causing
increased cell proliferation and decreased apoptosis, via the MAPK,
STAT5 and PI3K pathways (Hayakawa et al., 2000; Moore, Faisal,
et al., 2020; Papaemmanuil et al., 2016; The Cancer Genome
Atlas, 2013). Therefore, many different FLT3 inhibitors are at varying
stages in development to overcome the effects of this mutation.
Examples include gilteritinib, crenolanib and midostaurin (Aikawa
et al., 2020; McMahon et al., 2019; Piloto et al., 2007; Zhang
et al., 2019). These TKIs bind to the active conformation of FLT3 and
are at varying stages of approval. Gilteritinib and midostaurin are
FDA-approved while crenolanib is in Phase II trials (Galanis
2852 HEALY ET AL.
14765381,
2022,
12,
Downloaded
from
https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420,
Wiley
Online
Library
on
[11/06/2023].
See
the
Terms
and
Conditions
(https://onlinelibrary.wiley.com/terms-and-conditions)
on
Wiley
Online
Library
for
rules
of
use;
OA
articles
are
governed
by
the
applicable
Creative
Commons
License
et al., 2014; Levis, 2017; Levis et al., 2011; Zhang et al., 2019).
Although results for these drugs have all been promising, subsets of
patients exhibit resistance. This has, in part, been attributed to Ras
mutations, reactivating the MAPK and PI3K–AKT pathways
(McMahon et al., 2019; Zhang et al., 2019). These mutations were
detected in over 30% of patients who developed resistance to
gilteritinib, with Ras variant allele frequencies also increasing post-
drug exposure in patients who responded poorly to crenolanib. Inter-
estingly, not all resistant patients had FLT3 mutations following treat-
ment either, with different mutational signatures present instead
(McMahon et al., 2019; Zhang et al., 2019). This implies a clonal selec-
tion mechanism of resistance—a minor subpopulation at diagnosis
which became dominant following treatment and elimination of the
initial tumour burden. Taken together, it seems likely these Ras
mutations, either pre-existing or de novo, may contribute to resistance
to FLT3 inhibitors, through restoration of the original disease pheno-
type by expansion of an originally minor subclone. This is perhaps
unsurprising in acute myeloid leukaemia, which arises as a result of
clonal haematopoiesis (Desai et al., 2018).
7 | OPTIONS FOR TARGETING THE Ras
PATHWAY
With improved capability of detecting minor cancer subclones,
alongside the greater understanding of the impact of Ras mutations in
various cancers, the next considerable challenge is improved pharma-
cological targeting of Ras. However, drugging Ras has proven excep-
tionally difficult. A key drawback is the lack of available binding sites
for small molecule inhibitors. The nucleotide binding site (where GTP
or GDP binds) seems a desirable pocket to target, however the
picomolar affinity with which both GDP and GTP bind, as well as their
high intracellular concentrations, effectively outcompetes the binding
of any drug at this site (Cox et al., 2014; McCormick, 2018).
Therefore, targeting alternative proteins within Ras-regulated
pathways has been strongly investigated, with positive results seen,
for example with trametinib, the MEK inhibitor. Currently approved
for patients with BRAF V600-mutant metastatic melanoma or BRAF-
mutated (V600) non-small cell lung cancer, trametinib is well tolerated
in patients (Lugowska et al., 2015; Odogwu et al., 2018) and is also
being assessed in Ras-mutant myeloid malignancies (Borthakur
et al., 2016). This compound binds to phosphorylated MEK and
inhibits its downstream effectors (e.g. ERK), despite the presence of
constitutive Ras signalling. Subsequently, aberrant growth signalling
and apoptosis inhibition is reduced (Hofmann et al., 2012). Whilst this
has shown promising results (Borthakur et al., 2016; Lugowska
et al., 2015; Odogwu et al., 2018), this compound only inhibits the
MAPK pathway downstream of MEK, so constitutive activation of
other Ras-dependent pathways (e.g. PI3K–AKT) will still occur in the
presence of Ras mutations even when treated with this drug. In this
way, cancer can persist (Jones et al., 2019; Stinchcombe & John-
son, 2014). However, if Ras were to be targeted directly, signalling of
both of these pathways would be inhibited, leading to cell death.
As often seen with many diseases, combination of trametinib with
other therapeutics to inhibit multiple pathways together may reduce
the likelihood of continued cancer signalling and potential develop-
ment of resistance to this and other drugs (Infante & Swanton, 2014;
Planchard et al., 2016; Zhou et al., 2020). However, even with the
approved combination regimen of dabrafenib (BRAF inhibitor) with
trametinib (Lugowska et al., 2015), only the MAPK pathway is
inhibited, thereby maintaining the potential for aberrant PI3K path-
way signalling, which can in itself cause resistance to MEK inhibitors
(Jaiswal et al., 2009; Sos et al., 2009; Vitiello et al., 2019).
Alternatively, positive effects have been shown in vitro and in vivo
of co-administering AKT inhibitors with dabrafenib to inhibit two
Ras-mediated pathways (Lassen et al., 2014). However, combination
of the two classes of drugs in patients did not yield significant clinical
activity in reducing resistance seen in trametinib monotherapy, with
25% of patients exhibiting grade 3–4 toxicity (Algazi et al., 2018).
Taken together, whilst downstream pathway inhibition has proved
successful, this treatment method does have considerable disadvan-
tages and may not be a long-term solution for many patients.
Therefore, there is a distinct clinical need for novel means of treating
Ras-mutant cancers, which could include the targeting of Ras itself.
Given the difficulties with targeting Ras, inhibition of elements
upstream of Ras has been investigated. This includes inhibition of
GEFs including SOS Ras/Rac guanine nucleotide exchange factor 1 (
SOS1), to reduce the likelihood of Ras maintaining its GTP-bound
state and therefore inhibiting constitutive signalling (Evelyn
et al., 2014; Hillig et al., 2019). For example, BAY293 is a first-in-class
compound with the ability to bind directly to SOS1 and inhibit the
Ras–SOS interaction and thereby downstream signalling of the PI3K
and MAPK pathways (Hillig et al., 2019). Although the in vivo bioavail-
ability for this compound was poor, the concept of Ras–SOS inhibition
could prove useful in the future, with promising high throughput in
silico and in vitro screening results serving as a proof of concept for
inhibition of Ras via this mechanism (Evelyn et al., 2014, 2015; Hillig
et al., 2019). More recently, BI-1701963, a SOS1–pan-Ras interaction
inhibitor, has reached Phase I clinical trials, the first of its kind to do
so. Modified from the structure of BI-3406, a quinazoline-derived
compound, this novel inhibitor binds to the catalytic site of SOS1,
preventing its interaction with inactive KRAS, thus inhibiting activa-
tion (Gerlach et al., 2020; Hofmann et al., 2020).
In addition, alternative, more indirect pathways are also being
targeted as a means of inhibiting Ras, which may also be able to elimi-
nate the cancer stem cell. For example, the small molecule KYA1797K
has been shown to be effective against oncogenic Ras in colorectal
cancer and erlotinib-resistant non-small cell lung cancer (Park
et al., 2019). This compound indirectly targets Ras, through inhibition
of the Wnt/β-catenin pathway, which usually stabilises Ras (Jeong
et al., 2012; Moon et al., 2014). This pathway is also up-regulated in
cancer stem cells (Malanchi et al., 2008). KYA1797K has initiated anti-
tumour effects in KRAS-mutant cell lines and a KRAS-mutated mouse
model. KYA1797K also exhibited synergy with the current first-line
therapeutic regimen (cisplatin and pemetrexed) in non-small cell lung
cancer in vitro models (Park et al., 2019). However, KYA1797K
HEALY ET AL. 2853
14765381,
2022,
12,
Downloaded
from
https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420,
Wiley
Online
Library
on
[11/06/2023].
See
the
Terms
and
Conditions
(https://onlinelibrary.wiley.com/terms-and-conditions)
on
Wiley
Online
Library
for
rules
of
use;
OA
articles
are
governed
by
the
applicable
Creative
Commons
License
promoted apoptosis of both KRAS wild-type and mutant cells,
questioning the specificity of the drug for cancer cells whilst sparing
healthy cells. This study found both KRAS and β-catenin were
overexpressed in tumour regions compared to non-tumour regions
(Park et al., 2019). This may explain the limited toxicity seen during
KYA1797K studies, with low doses of KYA1797K having a more sub-
stantial effect on tumour cells, compared to healthy cells. Indeed, only
limited toxicity was seen during in vivo studies (Park et al., 2019).
However, protein expression varies considerably between tissues
(with KRAS particularly highly expressed in the brain) and so targeting
based on comparative expression between cancer and non-cancer
regions in one tissue may not represent overall toxicity potential
(Newlaczyl et al., 2017). Taken together, there may be a role for
KYA1797K as a concomitant therapy in non-small cell lung cancer
(as well as other cancers such as colorectal cancer). It presents a
means of eliminating both the primary cause of the disease (if KRAS-
mutated) and also minor subclones and cancer stem cells that could
give rise to resistance (Cho et al., 2020). Nevertheless, better under-
standing of the drug's effects on other tissues with high Ras expres-
sion must be gained.
8 | OPTIONS FOR TARGETING Ras—
DIRECT Ras TARGETING
Ras post-translational modifications were targeted as a means of
preventing Ras trafficking to the membrane and therefore
inhibiting downstream signalling. This included generation of
farnesyltransferase inhibitors, including lonafarnib and tipifarnib (Van
Cutsem et al., 2004). However, the effectiveness of these was
questionable, with most patients with KRAS-mutated diseases (such as
pancreatic cancer and leukaemia patients) receiving no clinical benefit
from these farnesyltransferase inhibitors (Borthakur et al., 2006;
Burnett et al., 2012; Harousseau et al., 2009; Van Cutsem et al., 2004).
Inefficacy was largely due to the redundancy mechanism of
geranylgeranyltransferase, which sufficiently modifies KRAS in the
absence of farnesyltransferase to permit its trafficking to the membrane
(Basso et al., 2006; Whyte et al., 1997). Interest in this strategy has
recently been revived with new personalised medicine approaches now
capable of identifying patients harbouring HRAS- or NRAS-driven can-
cers that are more likely to respond (Gilardi et al., 2020; Lee et al., 2020).
In recent years however, more promising steps have been made
regarding direct inhibition of oncogenic Ras. A key feature of this has
been the discovery of novel potential binding pockets for small mole-
cule inhibitors, to inhibit GEF activity or effector binding (Cruz-Migoni
et al., 2019; Maurer et al., 2012; Ostrem et al., 2013). Fragment-based
screening identified a previously undiscovered hydrophobic pocket
located between the Switch I and II regions (termed S-IIP), which was
successfully targeted by Ostrem et al. (2013) (Figure 4a). Binding of
peptide fragments was specific to G12C-mutated KRAS since the
compounds functioned through irreversible cysteine binding in this
particular pocket (and not with other cysteines found in wild-type
KRAS). Other key residues within this pocket include, but are not
limited to, V7, V9, M72, F78, Q99 and I100. In vitro models of KRAS
G12C-mutated lung cancer treated showed decreased survival upon
treatment with these compounds, with inactive Ras (Ras-GDP) levels
considerably greater than Ras-GTP. Further analysis showed that the
conformational disruption caused by binding of these fragments
reduced interactions with both SOS and effector molecules and
pathways, including B-RAF, C-RAF and the PI3K pathway (Gentile
et al., 2017; Ostrem et al., 2013).
The binding of these fragments to Ras also reduce SOS-catalysed
nucleotide exchange, a method of Ras inhibition which had been
previously explored. Compounds acting in this way either inhibited
conversion of Ras-GDP to Ras-GTP (Patgiri et al., 2011) or increased
the amount of Ras-GTP to such a level that it inhibited ERK phosphor-
ylation, since overactivation of Ras can be cytotoxic (the Ras “sweet-
spot model”) (Li et al., 2018). Either way, these compounds were
shown to inhibit the MAPK pathway but were largely tested against
Ras wild type (in the context of inhibiting the effects of RTK muta-
tions). Thus, the aforementioned compounds identified by Ostrem
et al. were revolutionary in their specificity for targeting Ras-mutated
disease.
Discovery and characterisation of this SII-P pocket have led to
the development of revolutionary KRAS G12C-selective covalent
inhibitors, including ARS-853 and latterly AMG-510 (Figure 4b–d)
(Canon et al., 2019; Patricelli et al., 2016). ARS-853 binds irreversibly
to the inactive form of KRAS G12C, preventing exchange of GDP for
GTP and therefore activation. This in turn inhibits downstream MAPK
and PI3K–AKT pathway signalling, with KRAS–CRAF interactions sig-
nificantly reduced. Moreover, in vitro evidence showed increased apo-
ptosis and cell cycle arrest in some (but not all) models tested (Lito
et al., 2016; Patricelli et al., 2016). This compound has subsequently
been fully characterised in silico and these studies revealed a dynamic
nature of the SII-P pocket, a feature which could be utilised in further
study (Khrenova et al., 2020).
Based on this work, alternative iterations of KRAS G12C inhibi-
tors have been produced. This was required since the probability of
ARS-853 locking KRAS in its inactive state in vivo was debatable,
given a lack of understanding regarding the cycling efficiency of Ras
between its inactive and active states. It had been deduced in vitro
that the G12C mutation permits rapid cycling of KRAS between these
states, hence permitting the binding of ARS-853 (Figure 4b). How-
ever, the possibility of finding the correct therapeutic window to
translate this compound to in vivo work proved complex (Janes
et al., 2018; Lito et al., 2016; Patricelli et al., 2016). Thus, alternatives
including ARS-1620 were developed to improve in vivo capability
(Figure 4c). Modifications to the ARS-853 structure resulted in
favourable pharmacokinetic (PK) properties, permitting a greater
understanding of KRAS activation status and dependency in vivo
(Janes et al., 2018). ARS-1620 is an orally bioavailable quinazoline-
based compound with limited side effects witnessed in preclinical
animal models (Janes et al., 2018; Li et al., 2018). Optimisation from
ARS-853 by inclusion of a fluorophenol, hydrophobic binding group
permitted stronger covalent (irreversible) binding within the SII-P
pocket (more specifically, interaction with H95 in this pocket), thus
2854 HEALY ET AL.
14765381,
2022,
12,
Downloaded
from
https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420,
Wiley
Online
Library
on
[11/06/2023].
See
the
Terms
and
Conditions
(https://onlinelibrary.wiley.com/terms-and-conditions)
on
Wiley
Online
Library
for
rules
of
use;
OA
articles
are
governed
by
the
applicable
Creative
Commons
License
FIGURE 4 Evolution of novel direct RAS-targeting agents. Chemical structure and protein structures of direct RAS-targeting agents in
complex with GDP-bound (pink) KRAS. (a) Minor modification of initial compound hit 6H05, 6H05 compound 6 (purple) bound covalently to
KRAS G12C, PDB accession no. 4LUC. (b) ARS-853 (purple) bound covalently to KRAS G12C (orange), PDB accession no. 5F2E. (c) ARS-1620
(purple) bound covalently to KRAS G12C (orange), PDB accession no. 5V9U. (d) AMG-510 (purple) bound to KRAS G12C (orange), PDB accession
no. 6OIM. (e) MRTX849 (purple) bound covalently to KRAS G12C (orange), PDB accession no. 6UT0. Molecules shown in relation to the switch
regions, largely binding in SII-P. All compounds bind covalently near to mutational hotspot
HEALY ET AL. 2855
14765381,
2022,
12,
Downloaded
from
https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420,
Wiley
Online
Library
on
[11/06/2023].
See
the
Terms
and
Conditions
(https://onlinelibrary.wiley.com/terms-and-conditions)
on
Wiley
Online
Library
for
rules
of
use;
OA
articles
are
governed
by
the
applicable
Creative
Commons
License
improving potency (Figure 4). The effects of this compound remained
mutation specific, thereby eliminating the risk of binding to KRAS wild
type in non-tumour cells and thus reduced toxicity potential. This
compound was well tolerated in patient-derived xenograft mice,
where a reduction in tumour burden through decreased Ras-mediated
downstream signalling was evident. This was the first example of an
in vivo trial using a compound targeting the SII-P pocket (Janes
et al., 2018).
From this, AMG-510 was developed, following further
modifications to the ARS-1620 structure, including addition of more
aromatic entities (Figure 4d). This enables AMG-510 to bind within a
slightly different groove of SII-P, enhancing potency and selectivity
(Canon et al., 2019). Promising preclinical in vitro and in vivo
experiments indicated arrest of the MAPK pathway and induction of a
pro-inflammatory tumour micro-environment. This compound has
since progressed into clinical trials, the first KRAS G12C-selective
inhibitor to do so. Early data indicate that out of 29 patients evaluable
for response at the time of publication, 5 exhibited partial response,
18 had stable disease and 6 had progressive disease. As with the
previous ARS-1620 animal studies, AMG-510 was generally well
tolerated, with no dose limiting toxicities recorded (Govindan, 2019;
Romero, 2020). However, larger cohorts and longer trials are
imperative in determining the true impact of this compound. Such
compounds could help combat the KRAS-mediated resistance to
monoclonal antibodies seen in non-small cell lung cancer and colorec-
tal cancer (Lièvre & Laurent-Puig, 2009; Park et al., 2019). However,
success of this compound and indeed this targeting mechanism, is
likely restricted to KRAS G12C-mutant cancer since some key
residues for binding of this compound are unique to KRAS and not
conserved between the different isoforms. Indeed, in vitro work
completed by Ostrem et al. (2013) illustrated that transduction of lung
cancer cell lines with KRAS G12V rescued the cancerous phenotype
(resistance to cell death, increased proliferation), thereby illustrating
how this mutation renders resistance to KRAS G12C inhibitors, as
expected.
Alternative KRAS G12C inhibitors are also in development and
showing considerable promise, including MRTX849 (Fell et al., 2020;
Hallin et al., 2020a). In a similar way to AMG-510 and the ARS com-
pounds discussed above, MRTX849 covalently binds to the inactive
form of KRAS, in SII-P (Figure 4e). This induces apoptosis through
down-regulation of the MAPK pathway. Interestingly, the PI3K–AKT
pathway remained relatively unaffected by MRTX849 (Hallin
et al., 2020b). In vivo trials with MRTX849 exhibited favourable phar-
macokinetic/pharmacodynamic properties (Fell et al., 2020) and both
cell line- and patient-derived xenograft modelling of pancreatic and
lung cancers also indicated up to a 30% reduction in tumour burden
(Fell et al., 2020; Hallin et al., 2020b). Individual patient case studies
from Phase I trials have also shown MRTX849 to be effective in
reducing tumour burden in both lung cancer and colorectal cancer,
although these data are largely incomplete (Hallin et al., 2020b). Taken
together, it is clear that KRAS G12C inhibitors have promise as a
means of abrogating Ras-mediated resistance, although there remain
drawbacks which need assessing, most notably, the lack of efficacy
against other Ras mutations, which are prevalent across Ras-mutated
disease.
Clearer understanding of the structure of Ras has permitted not
only elucidation of the SII-P pocket but also alternative binding sites,
including a pocket between the Switch I and II regions of Ras (termed
pocket I). Key residues available for interaction with small molecule
compounds include K5, L6, V7, V8, S39, D54, I55, L56, Y71, T74, G75
and E76 (Maurer et al., 2012; Quevedo et al., 2018). Antibody-frag-
ment-directed site exploration can be used to explore and exploit pre-
viously unconsidered drug interaction sites. In the case of Ras, this
mechanism has been used to analyse potential compound binding
sites within the previously identified pocket I that could be targeted
using small molecule inhibitors to interrupt effector proteins (such as
c-RAF and p110α) from binding to Ras via the Ras binding domain
(Maurer et al., 2012; Quevedo et al., 2018). This would provide an
alternative mechanism of abrogating the effects of oncogenic Ras
activation to those discussed previously and early results showed
effectiveness of antibody-derived compounds against a range of Ras
mutations and isoforms (Quevedo et al., 2018). However, given high
affinity binding of certain effectors to Ras (such as PI3K and B-RAF,
with 3.2- and 0.04-μM affinity, respectively) (Erijman &
Shifman, 2016), the high EC50s of the compounds identified in these
in vitro assays mean that many further modifications would be
required to convert these putative compounds into usable therapeu-
tics. Nevertheless, such antibody-derived fragments have the poten-
tial to be fused with small molecule protein–protein interaction
inhibitors to improve efficacy (Cruz-Migoni et al., 2019). Whilst many
Ras–effector interaction inhibitors have been trialled preclinically,
none have been implemented in the clinic in the context of Ras-
mutant cancer, owing to lack of efficacy or toxicity potential (Canon
et al., 2019; Keeton et al., 2017). However, crystal structure
determination showed that fusion of compounds developed through
antibody-derived fragment screening with known small molecule
protein–protein interaction inhibitors results in better binding within
pocket I, thus inhibiting Ras–effector interactions with a lower EC50.
Nevertheless, the therapeutic use of pocket I may be restricted since
such a pocket has also been detected in wild-type Ras (Cruz-Migoni
et al., 2019), thus increasing the risk for on-target toxicity.
In recent times, inhibition of the Ras–effector interaction has
been seen through competitive binding of rigosertib at the Ras bind-
ing domain. This compound elicits effects against MAPK, PI3K and
RAL pathway activation, in both wild-type and mutant Ras situations
(Athuluri-Divakar et al., 2016). Inhibition of multiple Ras-mediated
diseases have been seen in response to rigosertib, including pancre-
atic cancer and leukaemia (Athuluri-Divakar et al., 2016; Baker
et al., 2019). Rigosertib is moderately to well tolerated in clinical trials
thus far, although is yet to be specifically tested in the context of
Ras-mutant cancer (Bowles et al., 2014; Ma et al., 2012; Navada
et al., 2020). As with the antibody-derived, small molecule protein–
protein interaction inhibitors described above, toxicity may be as a
result of rigosertib's ability to target both wild-type and mutant KRAS
and thus, further studies are needed to fully assess the impact of such
a drug on healthy cells.
2856 HEALY ET AL.
14765381,
2022,
12,
Downloaded
from
https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420,
Wiley
Online
Library
on
[11/06/2023].
See
the
Terms
and
Conditions
(https://onlinelibrary.wiley.com/terms-and-conditions)
on
Wiley
Online
Library
for
rules
of
use;
OA
articles
are
governed
by
the
applicable
Creative
Commons
License
Antibody therapy is also currently being explored as means of
direct Ras targeting. However, a major drawback has been the capa-
bility of the antibody to cross the cell membrane, as with any protein-
based therapy (Bolhassani et al., 2017). Therefore, the development
of inRas37, a pan-Ras-targeting antibody, is a considerable step for-
ward in targeting Ras. Although the cellular uptake remains low
(approximately 4%), in vitro and in vivo work has shown promise in the
potential of inRas37 to inhibit both the MAPK and PI3K–AKT path-
ways, in a dose-dependent manner (Shin et al., 2020). Briefly, this drug
binds to integrins αVβ3 and αVβ5 on the cancer cell surface which
then undergo endocytosis. The antibody “escapes” the endosome as a
result of pH-determined cleavage (the antibody is cleaved from the
integrins better at pH 7, the cytoplasmic pH, compared to pH 6.5, the
pH of early endosomes) (Podinovskaia & Spang, 2018; Putnam, 2012;
Shin et al., 2020). The antibody then co-localises with Ras to block the
effector binding site, in a manner similar to the protein–protein inter-
action inhibitors described above. Mutations introduced into the gen-
eral antibody structure render it specific for mutant Ras binding, with
little activity against Ras wild type, thereby limiting on-target toxicity.
inRas37 has greater effect on cells with a greater dependency on Ras
signalling, which can include some tumour cells (Weinstein, 2002; Yi
et al., 2020). Nevertheless, when tested on large tumour models
(spheroids and cell-derived xenograft mice), efficacy decreased con-
siderably (Shin et al., 2020). This therefore shows that treatment with
this drug may be suitable for patients at earlier stages of Ras-mutated
cancer, before tumour burden is too great and reduces drug efficacy.
Thus, it could be used to treat Ras-initiated relapse as soon as it
occurs. However, its utility in eliminating minor subclones prior to
relapse is limited, given the low cellular uptake and the need for a high
Ras dependency in the cell.
9 | RESISTANCE TO Ras-TARGETING
AGENTS
Of course, there is potential for resistance to any therapeutic and
Ras-targeting drugs are no different. Some of these have been previ-
ously discussed here, such as the use of geranylgeranyltransferase to
overcome the effects of farnesyltransferase inhibitors (Whyte
et al., 1997) or the up-regulation of alternative Ras-mediated path-
ways, as seen when patients are treated with MEK inhibitors (Vitiello
et al., 2019). Other mechanisms of resistance are also possible, such
as the reactivation of ERK, which may be a potential resistance
mechanism in the case of a Ras-targeting agent (Bruner et al., 2017;
Ercan et al., 2012; Ochi et al., 2014). This has already been seen in the
case of EGFR-inhibitor resistance, whereby negative regulators of
ERK are down-regulated, so pro-apoptotic BIM is not fully up-
regulated and so cannot fully induce apoptosis. Alternatively, the gene
encoding ERK1, MAPK1, is amplified. These scenarios resulted in
in vitro and in vivo resistance to the putative EGFR inhibitor WZ4002
(Ercan et al., 2012). ERK reactivation has also been found to contrib-
ute to gefitinib resistance but, in this case, was found to be mediated
by Src (Ochi et al., 2014). Given these kinases are either side of Ras,
their co-operation could result in resistance to a Ras inhibitor. How-
ever, Src-mediated ERK reactivation is avoidable through treatment
with Src inhibitors (Ochi et al., 2014), which may present a means of
overcoming this potential Ras-inhibitor resistance mechanism. Taken
together, these studies imply that whilst resistance to Ras-targeting
drugs is possible, there are already means of overcoming this
resistance, just as a Ras-targeting drug would provide the means of
overcoming Ras-mediated resistance.
Whilst other elements of pathways contributing to resistance can
be targeted relatively easily, acquisition of secondary mutations in Ras
present a more pressing problem. For example, at present, the G12C-
specific inhibitor is perhaps the most developed means of inhibiting
Ras; however, a subsequent mutation in Ras would most likely
render this inhibitor ineffective. This has already been evidenced in
studies into KRAS G12C inhibitors, whereby rescue experiments
with the G12V mutation restored the cancerous phenotype (Ostrem
et al., 2013). Therefore, a pan-mutation-targeting drug, such as a
derivative of inRas37, may be favoured.
10 | DISCUSSION
Ras mutations in cancer and chemoresistance are important when
considering patient prognosis. Whilst it seems inevitable that resis-
tance will be an issue for a long time to come, better targeting of
potential causes is imperative. Advances in Ras inhibition could help
reduce the risk of resistance and relapse for a wide range of cancers,
given its mutational frequency. At present, some success has been
seen when multiple drugs are used to target different pathways impli-
cated in Ras-mutated disease. However, other studies have shown
lack of long-term efficacy when combining multiple therapies. There-
fore, single agent, multi-pathway targeting agents, including direct Ras
inhibitors, are becoming more heavily researched. It will be interesting
to see the effects of these in vivo, since this type of therapy may
reduce the potential for future development of drug resistance by
up-regulation of an alternative pathway. Targeting a common factor
at the centre of multiple pathways may target more cancer cell types
and thus reduce the heterogeneity of the tumour, eliminating
potential resistance causes before they become dominant.
Previous failures of Ras-targeting agents, including farnesyl-
transferase inhibitors, as well as perceived unfavourable protein
dynamics, have resulted in Ras being largely considered undruggable.
However, development of structural analysis techniques and a clearer
understanding of the key residues in Ras has altered this thinking, with
new compounds with novel, allele-specific Ras binding mechanisms
showing great promise. Phase I trials of AMG-510, a first-in-class direct
Ras-targeting agent, suggest a turning point has been reached in this
field of study. Whilst there is much more to be done, these preliminary
data indicate a solution for Ras-mediated resistance may be possible.
Nevertheless, a greater understanding of resistance must be
gained before relapse risk can be eliminated. Despite evidence
supporting the cancer stem cell theory, limited standard-of-care
detection sensitivities for initial diagnostic samples prevent
HEALY ET AL. 2857
14765381,
2022,
12,
Downloaded
from
https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420,
Wiley
Online
Library
on
[11/06/2023].
See
the
Terms
and
Conditions
(https://onlinelibrary.wiley.com/terms-and-conditions)
on
Wiley
Online
Library
for
rules
of
use;
OA
articles
are
governed
by
the
applicable
Creative
Commons
License
identification of the minor subclones present at diagnosis (McMahon
et al., 2019). It can therefore be difficult to determine likely causes of
relapse upon initial diagnosis and so constant monitoring for changes
in expression of genes commonly implicated in drug resistance, such
as Ras, may provide a useful tool for predicting disease trajectory.
Nevertheless, this is only useful if the effects of the acquired/emer-
gent mutations can be abrogated. In the cases discussed here,
improved analytical tools would ideally be combined with Ras-
targeting agents to prevent resistance taking hold.
Ultimately, chemoresistance, either intrinsic or acquired, due to
Ras mutations, whether primary or secondary, remains a considerable
problem. The concepts presented here, amongst many other examples,
illustrate the necessity for Ras-targeting drugs. There is a distinct clini-
cal requirement for the improved targeting of Ras in cancer, with Ras
implicated in both initial disease presentation and relapse. Although no
universal, direct Ras inhibitor has yet been achieved, considerable pro-
gress has been made in the last decade with the advent of allele-specific
inhibitors. This brings promise to the field, with the potential for better
treatment of Ras-initiated resistance a real prospect.
10.1 | Nomenclature of targets and ligands
Key protein targets and ligands in this article are hyperlinked to
corresponding entries in http://www.guidetopharmacology.org and
are permanently archived in the Concise Guide to PHARMACOLOGY
2019/20 (Alexander, Kelly, Mathie, Peters, et al., 2019; Alexander,
Fabbro, Kelly, Marrion, et al., 2019; Alexander, Fabbro, Kelly, Mathie,
et al., 2019).
CONFLICT OF INTEREST
The authors declare no conflicts of interest.
REFERENCES
Adnane, L., Trail, P. A., Taylor, I., & Wilhelm, S. M. (2006). Sorafenib
(BAY 43-9006, Nexavar®), a dual-action inhibitor that targets
RAF/MEK/ERK pathway in tumor cells and tyrosine kinases
VEGFR/PDGFR in tumor vasculature. In Methods in enzymology
(pp. 597–612). Academic Press.
Aikawa, T., Togashi, N., Iwanaga, K., Okada, H., Nishiya, Y., Inoue, S.,
Levis, M. J., & Isoyama, T. (2020). Quizartinib, a selective FLT3
inhibitor, maintains antileukemic activity in preclinical models of
RAS-mediated midostaurin-resistant acute myeloid leukemia cells.
Oncotarget, 11, 943–955. https://doi.org/10.18632/oncotarget.
27489
Alexander, S. P., Kelly, E., Mathie, A., Peters, J. A., Veale, E. L.,
Armstrong, J. F., Faccenda, E., Harding, S. D., Pawson, A. J.,
Sharman, J. L., Southan, C., Peter Buneman, O., Cidlowski, J. A.,
Christopoulos, A., Davenport, A. P., Fabbro, D., Spedding, M.,
Striessnig, J., Davies, J. A., & Collaborators, GTP. (2019). The concise
guide to PHARMACOLOGY 2019/20: Introduction and Other Protein
Targets. British Journal of Pharmacology, 176, S1–S20. https://doi.org/
10.1111/bph.14747
Alexander, S. P. H., Fabbro, D., Kelly, E., Marrion, N. V., Mathie, A.,
Peters, J. A., Veale, E. L., Armstrong, J. F., Faccenda, E., Harding, S. D.,
Pawson, A. J., Sharman, J. L., Southan, C., Davies, J. A., &
Collaborators, GTP. (2019). The concise guide to pharmacology
2019/20: Catalytic receptors. British Journal of Pharmacology, 176,
S247–S296. https://doi.org/10.1111/bph.14751
Alexander, S. P. H., Fabbro, D., Kelly, E., Mathie, A., Peters, J. A.,
Veale, E. L., Armstrong, J. F., Faccenda, E., Harding, S. D.,
Pawson, A. J., Sharman, J. L., Southan, C., Davies, J. A., & GTP collabo-
rators. (2019). The Concise Guide to PHARMACOLOGY 2019/20:
Enzymes. British Journal of Pharmacology, 176, S297–S396. https://
doi.org/10.1111/bph.14752
Algazi, A. P., Esteve-Puig, R., Nosrati, A., Hinds, B., Hobbs-
Muthukumar, A., Nandoskar, P., Ortiz-Urda, S., Chapman, P. B., &
Daud, A. (2018). Dual MEK/AKT inhibition with trametinib and
GSK2141795 does not yield clinical benefit in metastatic NRAS-
mutant and wild-type melanoma. Pigment Cell & Melanoma Research,
31, 110–114. https://doi.org/10.1111/pcmr.12644
Al-Hajj, M., Wicha, M. S., Benito-Hernandez, A., Morrison, S. J., &
Clarke, M. F. (2003). Prospective identification of tumorigenic breast
cancer cells. Proceedings of the National Academy of Sciences, 100,
3983–3988. https://doi.org/10.1073/pnas.0530291100
Amado, R. G., Wolf, M., Peeters, M., Van Cutsem, E., Siena, S.,
Freeman, D. J., Juan, T., Sikorski, R., Suggs, S., Radinsky, R.,
Patterson, S. D., & Chang, D. D. (2008). Wild-type KRAS is required
for panitumumab efficacy in patients with metastatic colorectal can-
cer. Journal of Clinical Oncology, 26, 1626–1634. https://doi.org/10.
1200/JCO.2007.14.7116
Amirouchene-Angelozzi, N., Swanton, C., & Bardelli, A. (2017). Tumor evo-
lution as a therapeutic target. Cancer Discovery, 7, 805–817. https://
doi.org/10.1158/2159-8290.CD-17-0343
André, F., Ciruelos, E., Rubovszky, G., Campone, M., Loibl, S.,
Rugo, H. S., et al. (2019). Alpelisib for PIK3CA-mutated, hormone
receptor–positive advanced breast cancer. New England Journal of
Medicine, 380, 1929–1940. https://doi.org/10.1056/NEJMoa
1813904
Athuluri-Divakar, S. K., Vasquez-Del Carpio, R., Dutta, K., Baker Stacey, J.,
Cosenza Stephen, C., Basu, I., Gupta, Y. K., Reddy, M. V. R., Ueno, L.,
Hart, J. R., Vogt, P. K., Mulholland, D., Guha, C., Aggarwal, A. K., &
Reddy, E. P. (2016). A small molecule RAS-mimetic disrupts RAS asso-
ciation with effector proteins to block signaling. Cell, 165, 643–655.
https://doi.org/10.1016/j.cell.2016.03.045
Baker, S. J., Cosenza, S. C., Ramana Reddy, M. V., & Premkumar Reddy, E.
(2019). Rigosertib ameliorates the effects of oncogenic KRAS signaling
in a murine model of myeloproliferative neoplasia. Oncotarget, 10,
1932–1942. https://doi.org/10.18632/oncotarget.26735
Basak, P., Sadhukhan, P., Sarkar, P., & Sil, P. C. (2017). Perspectives of
the Nrf-2 signaling pathway in cancer progression and therapy.
Toxicology Reports, 4, 306–318. https://doi.org/10.1016/j.toxrep.
2017.06.002
Basso, A. D., Kirschmeier, P., & Bishop, W. R. (2006). Thematic review
series: Lipid posttranslational modifications. Farnesyl transferase inhib-
itors. Journal of Lipid Research, 47, 15–31. https://doi.org/10.1194/jlr.
R500012-JLR200
Bhaduri, A., Di Lullo, E., Jung, D., Müller, S., Crouch, E. E., Espinosa, C. S.,
Ozawa, T., Alvarado, B., Spatazza, J., Cadwell, C. R., Wilkins, G.,
Velmeshev, D., Liu, S. J., Malatesta, M., Andrews, M. G., Mostajo-
Radji, M. A., Huang, E. J., Nowakowski, T. J., Lim, D. A., …
Kriegstein, A. R. (2020). Outer radial glia-like cancer stem cells contrib-
ute to heterogeneity of glioblastoma. Cell Stem Cell, 26, 48–63.e46.
https://doi.org/10.1016/j.stem.2019.11.015
Bhagwat, S. V., McMillen, W. T., Cai, S., Zhao, B., Whitesell, M., Shen, W.,
Kindler, L., Flack, R. S., Wu, W., Anderson, B., Zhai, Y., Yuan, X.-J.,
Pogue, M., Van Horn, Rao, X., McCann, D., Dropsey, A. J., Manro, J.,
Walgren, J., … Peng, S.-B. (2020). ERK inhibitor LY3214996 targets
ERK pathway–driven cancers: A therapeutic approach toward preci-
sion medicine. Molecular Cancer Therapeutics, 19, 325–336. https://
doi.org/10.1158/1535-7163.MCT-19-0183
2858 HEALY ET AL.
14765381,
2022,
12,
Downloaded
from
https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420,
Wiley
Online
Library
on
[11/06/2023].
See
the
Terms
and
Conditions
(https://onlinelibrary.wiley.com/terms-and-conditions)
on
Wiley
Online
Library
for
rules
of
use;
OA
articles
are
governed
by
the
applicable
Creative
Commons
License
Bolhassani, A., Jafarzade, B. S., & Mardani, G. (2017). In vitro and in vivo
delivery of therapeutic proteins using cell penetrating peptides. Pep-
tides, 87, 50–63. https://doi.org/10.1016/j.peptides.2016.11.011
Bonnet, D., & Dick, J. E. (1997). Human acute myeloid leukemia is
organized as a hierarchy that originates from a primitive hematopoi-
etic cell. Nature Medicine, 3, 730–737. https://doi.org/10.1038/
nm0797-730
Borthakur, G., Kantarjian, H., Daley, G., Talpaz, M., O'Brien, S., Garcia-
Manero, G., Giles, F., Faderl, S., Sugrue, M., & Cortes, J. (2006). Pilot
study of lonafarnib, a farnesyl transferase inhibitor, in patients with
chronic myeloid leukemia in the chronic or accelerated phase that is
resistant or refractory to imatinib therapy. Cancer, 106, 346–352.
https://doi.org/10.1002/cncr.21590
Borthakur, G., Popplewell, L., Boyiadzis, M., Foran, J., Platzbecker, U.,
Vey, N., Walter, R. B., Olin, R., Raza, A., Giagounidis, A., Al-Kali, A.,
Jabbour, E., Kadia, T., Garcia-Manero, G., Bauman, J. W., Wu, Y.,
Liu, Y., Schramek, D., Cox, D. S., … Kantarjian, H. (2016). Activity of
the oral mitogen-activated protein kinase kinase inhibitor trametinib in
RAS-mutant relapsed or refractory myeloid malignancies. Cancer, 122,
1871–1879. https://doi.org/10.1002/cncr.29986
Bowles, D. W., Diamond, J. R., Lam, E. T., Weekes, C. D., Astling, D. P.,
Anderson, R. T., Leong, S., Gore, L., Varella-Garcia, M., Vogler, B. W.,
Keysar, S. B., Freas, E., Aisner, D. L., Ren, C., Tan, A.-C., Wilhelm, F.,
Maniar, M., Eckhardt, S. G., Messersmith, W. A., & Jimeno, A. (2014).
Phase I study of oral rigosertib (ON 01910.Na), a dual inhibitor of the
PI3K and Plk1 pathways, in adult patients with advanced solid malig-
nancies. Clinical Cancer Research, 20, 1656–1665. https://doi.org/10.
1158/1078-0432.CCR-13-2506
Bowyer, S., Lee, R., Fusi, A., & Lorigan, P. (2015). Dabrafenib and its use in
the treatment of metastatic melanoma. Melanoma Manag, 2, 199–208.
https://doi.org/10.2217/mmt.15.21
Boyer, L. A., Lee, T. I., Cole, M. F., Johnstone, S. E., Levine, S. S.,
Zucker, J. P., Guenther, M. G., Kumar, R. M., Murray, H. L.,
Jenner, R. G., Gifford, D. K., Melton, D. A., Jaenisch, R., & Young, R. A.
(2005). Core transcriptional regulatory circuitry in human embryonic
stem cells. Cell, 122, 947–956. https://doi.org/10.1016/j.cell.2005.
08.020
Bruner, J. K., Ma, H. S., Li, L., Qin, A. C. R., Rudek, M. A., Jones, R. J.,
Levis, M. J., Pratz, K. W., Pratilas, C. A., & Small, D. (2017). Adaptation
to TKI treatment reactivates ERK signaling in tyrosine kinase–driven
leukemias and other malignancies. Cancer Research, 77, 5554–5563.
https://doi.org/10.1158/0008-5472.CAN-16-2593
Burnett, A. K., Russell, N. H., Culligan, D., Cavanagh, J., Kell, J.,
Wheatley, K., Virchis, A., Hills, R. K., Milligan, D., & AML Working
Group of the UK National Cancer Research Institute. (2012). The addi-
tion of the farnesyl transferase inhibitor, tipifarnib, to low dose
cytarabine does not improve outcome for older patients with AML.
British Journal of Haematology, 158, 519–522. https://doi.org/10.
1111/j.1365-2141.2012.09165.x
Burrell, R. A., McGranahan, N., Bartek, J., & Swanton, C. (2013). The causes
and consequences of genetic heterogeneity in cancer evolution.
Nature, 501, 338–345. https://doi.org/10.1038/nature12625
Caiola, E., Salles, D., Frapolli, R., Lupi, M., Rotella, G., Ronchi, A.,
Garassino, M. C., Mattschas, N., Colavecchio, S., Broggini, M.,
Wiesmüller, L., & Marabese, M. (2015). Base excision repair-mediated
resistance to cisplatin in KRAS(G12C) mutant NSCLC cells. Oncotarget,
6, 30072–30087. https://doi.org/10.18632/oncotarget.5019
Canon, J., Rex, K., Saiki, A. Y., Mohr, C., Cooke, K., Bagal, D., Gaida, K.,
Holt, T., Knutson, C. G., Koppada, N., Lanman, B. A., Werner, J.,
Rapaport, A. S., Miguel, T. S., Ortiz, R., Osgood, T., Sun, J.-R., Zhu, X.,
McCarter, J. D., … Lipford, J. R. (2019). The clinical KRAS(G12C) inhibi-
tor AMG 510 drives anti-tumour immunity. Nature, 575, 217–223.
https://doi.org/10.1038/s41586-019-1694-1
Cantley, L. C. (2002). The phosphoinositide 3-kinase pathway. Science,
296, 1655–1657. https://doi.org/10.1126/science.296.5573.1655
Castellano, E., & Downward, J. (2011). RAS interaction with PI3K: More
than just another effector pathway. Genes & Cancer, 2, 261–274.
https://doi.org/10.1177/1947601911408079
Cerami, E., Gao, J., Dogrusoz, U., Gross, B. E., Sumer, S. O., Aksoy, B. A.,
Jacobsen, A., Byrne, C. J., Heuer, M. L., Larsson, E., Antipin, Y.,
Reva, B., Goldberg, A. P., Sander, C., & Schultz, N. (2012). The cBio
cancer genomics portal: An open platform for exploring multi-
dimensional cancer genomics data. Cancer Discovery, 2, 401–404.
https://doi.org/10.1158/2159-8290.CD-12-0095
Cha, P. H., Cho, Y. H., Lee, S. K., Lee, J., Jeong, W. J., Moon, B. S., Yun, J.-
H., Yang, J. S., Choi, S., Yoon, J., Kim, H.-Y., Kim, M.-Y., Kaduwal, S.,
Lee, W., Min, D. S., Kim, H., Han, G., & Choi, K.-Y. (2016). Small-
molecule binding of the axin RGS domain promotes β-catenin and Ras
degradation. Nature Chemical Biology, 12, 593–600. https://doi.org/
10.1038/nchembio.2103
Chan, L. H., Zhou, L., Ng, K. Y., Wong, T. L., Lee, T. K., Sharma, R.,
Loong, J. H., Ching, Y. P., Yuan, Y.-F., Xie, D., Lo, C. M., Man, K.,
Artegiani, B., Clevers, H., Yan, H. H., Leung, S. Y., Richard, S., Guan, X.-
Y., Huen, M. S. Y., & Ma, S. (2018). PRMT6 regulates RAS/RAF binding
and MEK/ERK-mediated cancer stemness activities in hepatocellular
carcinoma through CRAF methylation. Cell Reports, 25, 690–701e698.
https://doi.org/10.1016/j.celrep.2018.09.053
Chang, F., Lee, J. T., Navolanic, P. M., Steelman, L. S., Shelton, J. G.,
Blalock, W. L., Franklin, R. A., & McCubrey, J. A. (2003). Involvement
of PI3K/Akt pathway in cell cycle progression, apoptosis, and neoplas-
tic transformation: A target for cancer chemotherapy. Leukemia, 17,
590–603. https://doi.org/10.1038/sj.leu.2402824
Cho, Y.-H., Ro, E. J., Yoon, J.-S., Kwak, D.-K., Cho, J., Kang, D. W., Lee, H.-
Y., & Choi, K.-Y. (2020). Small molecule-induced simultaneous
destabilization of β-catenin and RAS is an effective molecular strategy
to suppress stemness of colorectal cancer cells. Cell Communication
and Signaling: CCS, 18, 38. https://doi.org/10.1186/s12964-020-
0519-z
Ciardiello, F., & Tortora, G. (2008). EGFR antagonists in cancer treatment.
New England Journal of Medicine, 358, 1160–1174. https://doi.org/10.
1056/NEJMra0707704
Corces-Zimmerman, M. R., Hong, W. J., Weissman, I. L.,
Medeiros, B. C., & Majeti, R. (2014). Preleukemic mutations in
human acute myeloid leukemia affect epigenetic regulators and per-
sist in remission. Proceedings of the National Academy of Sciences of
the United States of America, 111, 2548–2553. https://doi.org/10.
1073/pnas.1324297111
Cox, A. D., & Der, C. J. (2002). Farnesyltransferase inhibitors: Promises
and realities. Current Opinion in Pharmacology, 2, 388–393. https://doi.
org/10.1016/S1471-4892(02)00181-9
Cox, A. D., Fesik, S. W., Kimmelman, A. C., Luo, J., & Der, C. J. (2014).
Drugging the undruggable RAS: Mission possible? Nature Reviews Drug
Discovery, 13, 828–851. https://doi.org/10.1038/nrd4389
Cruz-Migoni, A., Canning, P., Quevedo, C. E., Bataille, C. J. R., Bery, N.,
Miller, A., Russell, A. J., Phillips, S. E. V., Carr, S. B., & Rabbitts, T. H.
(2019). Structure-based development of new RAS-effector inhibitors
from a combination of active and inactive RAS-binding compounds.
Proceedings of the National Academy of Sciences, 116, 2545–2550.
https://doi.org/10.1073/pnas.1811360116
D'Abaco, G. M., Whitehead, R. H., & Burgess, A. W. (1996). Synergy
between Apc min and an activated ras mutation is sufficient to induce
colon carcinomas. Molecular and Cellular Biology, 16, 884–891. https://
doi.org/10.1128/MCB.16.3.884
De Roock, Claes, B., Bernasconi, D., De Schutter, Biesmans, B.,
Fountzilas, G., Kalogeras, K. T., Kotoula, V., Papamichael, D., Laurent-
Puig, P., Penault-Llorca, F., Rougier, P., Vincenzi, B., Santini, D.,
Tonini, G., Cappuzzo, F., Frattini, M., Molinari, F., Saletti, P., …
Tejpar, S. (2010). Effects of KRAS, BRAF, NRAS, and PIK3CA muta-
tions on the efficacy of cetuximab plus chemotherapy in
chemotherapy-refractory metastatic colorectal cancer: A retrospective
HEALY ET AL. 2859
14765381,
2022,
12,
Downloaded
from
https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420,
Wiley
Online
Library
on
[11/06/2023].
See
the
Terms
and
Conditions
(https://onlinelibrary.wiley.com/terms-and-conditions)
on
Wiley
Online
Library
for
rules
of
use;
OA
articles
are
governed
by
the
applicable
Creative
Commons
License
consortium analysis. The Lancet Oncology, 11, 753–762. https://doi.
org/10.1016/S1470-2045(10)70130-3
de Vries, G., Rosas-Plaza, X., van Vugt, M. A. T. M., Gietema, J. A., & de
Jong, S. (2020). Testicular cancer: Determinants of cisplatin sensitivity
and novel therapeutic opportunities. Cancer Treatment Reviews, 88,
102054. https://doi.org/10.1016/j.ctrv.2020.102054
Dean, M., Fojo, T., & Bates, S. (2005). Tumour stem cells and drug resis-
tance. Nature Reviews Cancer, 5, 275–284. https://doi.org/10.1038/
nrc1590
DeNicola, G. M., Karreth, F. A., Humpton, T. J., Gopinathan, A., Wei, C.,
Frese, K., Mangal, D., Yu, K. H., Yeo, C. J., Calhoun, E. S., Scrimieri, F.,
Winter, J. M., Hruban, R. H., Iacobuzio-Donahue, C., Kern, S. E.,
Blair, I. A., & Tuveson, D. A. (2011). Oncogene-induced Nrf2 transcrip-
tion promotes ROS detoxification and tumorigenesis. Nature, 475,
106–109. https://doi.org/10.1038/nature10189
Desai, P., Mencia-Trinchant, N., Savenkov, O., Simon, M. S., Cheang, G.,
Lee, S., Samuel, M., Ritchie, E. K., Guzman, M. L., Ballman, K. V.,
Roboz, G. J., & Hassane, D. C. (2018). Somatic mutations precede
acute myeloid leukemia years before diagnosis. Nature Medicine, 24,
1015–1023. https://doi.org/10.1038/s41591-018-0081-z
Diaz, L. A. Jr., Williams, R. T., Wu, J., Kinde, I., Hecht, J. R., Berlin, J.,
Allen, B., Bozic, I., Reiter, J. G., Nowak, M. A., Kinzler, K. W.,
Oliner, K. S., & Vogelstein, B. (2012). The molecular evolution of
acquired resistance to targeted EGFR blockade in colorectal cancers.
Nature, 486, 537–540. https://doi.org/10.1038/nature11219
Du, R., Shen, W., Liu, Y., Gao, W., Zhou, W., Li, J., Zhao, S., Chen, C.,
Chen, Y., Liu, Y., Sun, P., Xiang, R., Shi, Y., & Luo, Y. (2019). TGIF2 pro-
motes the progression of lung adenocarcinoma by bridging
EGFR/RAS/ERK signaling to cancer cell stemness. Signal Transduction
and Targeted Therapy, 4, 60. https://doi.org/10.1038/s41392-019-
0098-x
Duffy, M. J., & Crown, J. (2021). Drugging “undruggable” genes for cancer
treatment: Are we making progress? International Journal of Cancer,
148, 8–17. https://doi.org/10.1002/ijc.33197
Dunn, C., & Croom, K. F. (2006). Everolimus. Drugs, 66, 547–570. https://
doi.org/10.2165/00003495-200666040-00009
Duronio, V. (2008). The life of a cell: Apoptosis regulation by the
PI3K/PKB pathway. Biochemical Journal, 415, 333–344. https://doi.
org/10.1042/BJ20081056
Eberlein, C. A., Stetson, D., Markovets, A. A., Al-Kadhimi, K. J., Lai, Z.,
Fisher, P. R., Meador, C. B., Spitzler, P., Ichihara, E., Ross, S. J.,
Ahdesmaki, M. J., Ahmed, A., Ratcliffe, L. E., Christey O'Brien, E. L.,
Barnes, C. H., Brown, H., Smith, P. D., Dry, J. R., Beran, G., …
Cross, D. A. E. (2015). Acquired resistance to the mutant-selective
EGFR inhibitor AZD9291 is associated with increased dependence on
RAS signaling in preclinical models. Cancer Research, 75, 2489–2500.
https://doi.org/10.1158/0008-5472.CAN-14-3167
Engelman, J. A., Luo, J., & Cantley, L. C. (2006). The evolution of pho-
sphatidylinositol 3-kinases as regulators of growth and metabolism.
Nature Reviews Genetics, 7, 606–619. https://doi.org/10.1038/
nrg1879
Ercan, D., Xu, C., Yanagita, M., Monast, C. S., Pratilas, C. A., Montero, J.,
Butaney, M., Shimamura, T., Sholl, L., Ivanova, E. V., Tadi, M.,
Rogers, A., Repellin, C., Capelletti, M., Maertens, O., Goetz, E. M.,
Letai, A., Garraway, L. A., Lazzara, M. J., … Jänne, P. A. (2012). Reac-
tivation of ERK signaling causes resistance to EGFR kinase inhibitors.
Cancer Discovery, 2, 934–947. https://doi.org/10.1158/2159-8290.
CD-12-0103
Erijman, A., & Shifman, J. M. (2016). RAS/effector interactions from
structural and biophysical perspective. Mini Reviews in Medicinal
Chemistry, 16, 370–375. https://doi.org/10.2174/
1389557515666151001141838
Evelyn, C. R., Biesiada, J., Duan, X., Tang, H., Shang, X., Papoian, R.,
Seibel, W. L., Nelson, S., Meller, J., & Zheng, Y. (2015). Combined ratio-
nal design and a high throughput screening platform for identifying
chemical inhibitors of a Ras-activating enzyme. Journal of Biological
Chemistry, 290, 12879–12898. https://doi.org/10.1074/jbc.M114.
634493
Evelyn, C. R., Duan, X., Biesiada, J., Seibel William, L., Meller, J., &
Zheng, Y. (2014). Rational design of small molecule inhibitors targeting
the Ras GEF, SOS1. Chemistry & Biology, 21, 1618–1628. https://doi.
org/10.1016/j.chembiol.2014.09.018
Feldman, D. R., Iyer, G., Van Alstine, L., Patil, S., Al-Ahmadie, H.,
Reuter, V. E., Bosl, G. J., Chaganti, R. S., & Solit, D. B. (2014).
Presence of somatic mutations within PIK3CA, AKT, RAS, and FGFR3
but not BRAF in cisplatin-resistant germ cell tumors. Clinical Cancer
Research, 20, 3712–3720. https://doi.org/10.1158/1078-0432.CCR-
13-2868
Fell, J. B., Fischer, J. P., Baer, B. R., Blake, J. F., Bouhana, K., Briere, D. M.,
Brown, K. D., Burgess, L. E., Burns, A. C., Burkard, M. R., Chiang, H.,
Chicarelli, M. J., Cook, A. W., Gaudino, J. J., Hallin, J., Hanson, L.,
Hartley, D. P., Hicken, E. J., Hingorani, G. P., … Marx, M. A. (2020).
Identification of the clinical development candidate MRTX849, a
covalent KRASG12C inhibitor for the treatment of cancer. Journal of
Medicinal Chemistry, 63, 6679–6693. https://doi.org/10.1021/acs.
jmedchem.9b02052
Ferino, A., Rapozzi, V., & Xodo, L. E. (2020). The ROS-KRAS-Nrf2 axis in
the control of the redox homeostasis and the intersection with
survival-apoptosis pathways: Implications for photodynamic therapy.
Journal of Photochemistry and Photobiology B: Biology, 202, 111672.
https://doi.org/10.1016/j.jphotobiol.2019.111672
Freeman, W. H. (2000). Section 20.4 receptor tyrosine kinases and RAS. In
H. B. A. Lodish, S. L. Zipursky, et al. (Eds.), Molecular cell biology (4th
ed.). New York: W.H. Freeman.
Furet, P., Guagnano, V., Fairhurst, R. A., Imbach-Weese, P., Bruce, I.,
Knapp, M., Fritsch, C., Blasco, F., Blanz, J., Aichholz, R., Hamon, J.,
Fabbro, D., & Caravatti, G. (2013). Discovery of NVP-BYL719 a potent
and selective phosphatidylinositol-3 kinase alpha inhibitor selected for
clinical evaluation. Bioorganic & Medicinal Chemistry Letters, 23,
3741–3748. https://doi.org/10.1016/j.bmcl.2013.05.007
Galanis, A., Ma, H., Rajkhowa, T., Ramachandran, A., Small, D., Cortes, J., &
Levis, M. (2014). Crenolanib is a potent inhibitor of FLT3 with activity
against resistance-conferring point mutants. Blood, 123, 94–100.
https://doi.org/10.1182/blood-2013-10-529313
Gao, J., Aksoy, B. A., Dogrusoz, U., Dresdner, G., Gross, B., Sumer, S. O.,
Sun, Y., Jacobsen, A., Sinha, R., Larsson, E., Cerami, E., Sander, C., &
Schultz, N. (2013). Integrative analysis of complex cancer genomics
and clinical profiles using the cBioPortal. Science Signaling, 6, pl1.
https://doi.org/10.1126/scisignal.2004088
Garassino, M. C., Marabese, M., Rusconi, P., Rulli, E., Martelli, O.,
Farina, G., Scanni, A., & Broggini, M. (2011). Different types of K-Ras
mutations could affect drug sensitivity and tumour behaviour in non-
small-cell lung cancer. Annals of Oncology, 22, 235–237. https://doi.
org/10.1093/annonc/mdq680
Garnock-Jones, K. P. (2015). Cobimetinib: First global approval. Drugs, 75,
1823–1830. https://doi.org/10.1007/s40265-015-0477-8
Gentile, D. R., Rathinaswamy, M. K., Jenkins, M. L., Moss, S. M.,
Siempelkamp, B. D., Renslo, A. R., Burke, J. E., & Shokat, K. M. (2017).
Ras binder induces a modified switch-II pocket in GTP and GDP states.
Cell Chemical Biology, 24, 1455–1466e1414. https://doi.org/10.1016/
j.chembiol.2017.08.025
Gerlach, D., Gmachl, M., Ramharter, J., Teh, J., Fu, S.-C., Trapani, F.,
Kessler, D., Rumpel, K., Botesteanu, D.-A., Ettmayer, P., Arnhof, H.,
Gerstberger, T., Kofink, C., Wunberg, T., Vellano, C. P.,
Heffernan, T. P., Marszalek, J. R., Pearson, M., McConnell, D. B., …
Hofmann, M. H. (2020). Abstract 1091: BI-3406 and BI 1701963:
Potent and selective SOS1::KRAS inhibitors induce regressions
in combination with MEK inhibitors or irinotecan. Cancer Research,
80, 1091.
2860 HEALY ET AL.
14765381,
2022,
12,
Downloaded
from
https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420,
Wiley
Online
Library
on
[11/06/2023].
See
the
Terms
and
Conditions
(https://onlinelibrary.wiley.com/terms-and-conditions)
on
Wiley
Online
Library
for
rules
of
use;
OA
articles
are
governed
by
the
applicable
Creative
Commons
License
British J Pharmacology - 2021 - Healy - The importance of Ras in drug resistance in cancer.pdf
British J Pharmacology - 2021 - Healy - The importance of Ras in drug resistance in cancer.pdf
British J Pharmacology - 2021 - Healy - The importance of Ras in drug resistance in cancer.pdf
British J Pharmacology - 2021 - Healy - The importance of Ras in drug resistance in cancer.pdf
British J Pharmacology - 2021 - Healy - The importance of Ras in drug resistance in cancer.pdf
British J Pharmacology - 2021 - Healy - The importance of Ras in drug resistance in cancer.pdf
British J Pharmacology - 2021 - Healy - The importance of Ras in drug resistance in cancer.pdf

More Related Content

Similar to British J Pharmacology - 2021 - Healy - The importance of Ras in drug resistance in cancer.pdf

Lancet Oncology 2004
Lancet Oncology 2004Lancet Oncology 2004
Lancet Oncology 2004
Ruth Foley
 
Mathur A Biochimie 2016
Mathur A Biochimie 2016Mathur A Biochimie 2016
Mathur A Biochimie 2016
Aditi Mathur
 
Cancer Medicine - 2023 - Ma - Novel strategies to reverse chemoresistance in ...
Cancer Medicine - 2023 - Ma - Novel strategies to reverse chemoresistance in ...Cancer Medicine - 2023 - Ma - Novel strategies to reverse chemoresistance in ...
Cancer Medicine - 2023 - Ma - Novel strategies to reverse chemoresistance in ...
damodara kumaran
 
Liver International 2011
Liver International 2011Liver International 2011
Liver International 2011
mbouattour
 
Mechanistic Studies and Modeling Reveal the Origin of Differential Inhibition...
Mechanistic Studies and Modeling Reveal the Origin of Differential Inhibition...Mechanistic Studies and Modeling Reveal the Origin of Differential Inhibition...
Mechanistic Studies and Modeling Reveal the Origin of Differential Inhibition...
Ira Dicker
 
Cell, Vol. 100, 57–70, January 7, 2000, Copyright 2000 by Cel.docx
Cell, Vol. 100, 57–70, January 7, 2000, Copyright 2000 by Cel.docxCell, Vol. 100, 57–70, January 7, 2000, Copyright 2000 by Cel.docx
Cell, Vol. 100, 57–70, January 7, 2000, Copyright 2000 by Cel.docx
zebadiahsummers
 
Cell, Vol. 100, 57–70, January 7, 2000, Copyright 2000 by Cel.docx
Cell, Vol. 100, 57–70, January 7, 2000, Copyright 2000 by Cel.docxCell, Vol. 100, 57–70, January 7, 2000, Copyright 2000 by Cel.docx
Cell, Vol. 100, 57–70, January 7, 2000, Copyright 2000 by Cel.docx
keturahhazelhurst
 

Similar to British J Pharmacology - 2021 - Healy - The importance of Ras in drug resistance in cancer.pdf (20)

Vol 87-no2-article-3
Vol 87-no2-article-3Vol 87-no2-article-3
Vol 87-no2-article-3
 
SCT60103 Group 2 Presentation
SCT60103 Group 2 PresentationSCT60103 Group 2 Presentation
SCT60103 Group 2 Presentation
 
Lancet Oncology 2004
Lancet Oncology 2004Lancet Oncology 2004
Lancet Oncology 2004
 
Immuotherapy
ImmuotherapyImmuotherapy
Immuotherapy
 
Mathur A Biochimie 2016
Mathur A Biochimie 2016Mathur A Biochimie 2016
Mathur A Biochimie 2016
 
Cancer Medicine - 2023 - Ma - Novel strategies to reverse chemoresistance in ...
Cancer Medicine - 2023 - Ma - Novel strategies to reverse chemoresistance in ...Cancer Medicine - 2023 - Ma - Novel strategies to reverse chemoresistance in ...
Cancer Medicine - 2023 - Ma - Novel strategies to reverse chemoresistance in ...
 
Terapia ayurvedica no cancer
Terapia ayurvedica no cancerTerapia ayurvedica no cancer
Terapia ayurvedica no cancer
 
How patient subpopulations are changing the commercialization of oncology pro...
How patient subpopulations are changing the commercialization of oncology pro...How patient subpopulations are changing the commercialization of oncology pro...
How patient subpopulations are changing the commercialization of oncology pro...
 
Immunotherapy
ImmunotherapyImmunotherapy
Immunotherapy
 
Clinic Correlation and Prognostic Value of P4HB and GRP78 Expression in Gastr...
Clinic Correlation and Prognostic Value of P4HB and GRP78 Expression in Gastr...Clinic Correlation and Prognostic Value of P4HB and GRP78 Expression in Gastr...
Clinic Correlation and Prognostic Value of P4HB and GRP78 Expression in Gastr...
 
Clinic Correlation and Prognostic Value of P4HB and GRP78 Expression in Gastr...
Clinic Correlation and Prognostic Value of P4HB and GRP78 Expression in Gastr...Clinic Correlation and Prognostic Value of P4HB and GRP78 Expression in Gastr...
Clinic Correlation and Prognostic Value of P4HB and GRP78 Expression in Gastr...
 
Clinic Correlation and Prognostic Value of P4HB and GRP78 Expression in Gastr...
Clinic Correlation and Prognostic Value of P4HB and GRP78 Expression in Gastr...Clinic Correlation and Prognostic Value of P4HB and GRP78 Expression in Gastr...
Clinic Correlation and Prognostic Value of P4HB and GRP78 Expression in Gastr...
 
Clinic Correlation and Prognostic Value of P4HB and GRP78 Expression in Gastr...
Clinic Correlation and Prognostic Value of P4HB and GRP78 Expression in Gastr...Clinic Correlation and Prognostic Value of P4HB and GRP78 Expression in Gastr...
Clinic Correlation and Prognostic Value of P4HB and GRP78 Expression in Gastr...
 
2000 weinberg. hallsmarcks of cancer new generation
2000 weinberg. hallsmarcks of cancer new generation2000 weinberg. hallsmarcks of cancer new generation
2000 weinberg. hallsmarcks of cancer new generation
 
Gene therapy for the treatmen of cancer
Gene therapy for the treatmen of cancerGene therapy for the treatmen of cancer
Gene therapy for the treatmen of cancer
 
Vital Signs Edition #3
Vital Signs   Edition #3Vital Signs   Edition #3
Vital Signs Edition #3
 
Liver International 2011
Liver International 2011Liver International 2011
Liver International 2011
 
Mechanistic Studies and Modeling Reveal the Origin of Differential Inhibition...
Mechanistic Studies and Modeling Reveal the Origin of Differential Inhibition...Mechanistic Studies and Modeling Reveal the Origin of Differential Inhibition...
Mechanistic Studies and Modeling Reveal the Origin of Differential Inhibition...
 
Cell, Vol. 100, 57–70, January 7, 2000, Copyright 2000 by Cel.docx
Cell, Vol. 100, 57–70, January 7, 2000, Copyright 2000 by Cel.docxCell, Vol. 100, 57–70, January 7, 2000, Copyright 2000 by Cel.docx
Cell, Vol. 100, 57–70, January 7, 2000, Copyright 2000 by Cel.docx
 
Cell, Vol. 100, 57–70, January 7, 2000, Copyright 2000 by Cel.docx
Cell, Vol. 100, 57–70, January 7, 2000, Copyright 2000 by Cel.docxCell, Vol. 100, 57–70, January 7, 2000, Copyright 2000 by Cel.docx
Cell, Vol. 100, 57–70, January 7, 2000, Copyright 2000 by Cel.docx
 

Recently uploaded

👉 Chennai Sexy Aunty’s WhatsApp Number 👉📞 7427069034 👉📞 Just📲 Call Ruhi Colle...
👉 Chennai Sexy Aunty’s WhatsApp Number 👉📞 7427069034 👉📞 Just📲 Call Ruhi Colle...👉 Chennai Sexy Aunty’s WhatsApp Number 👉📞 7427069034 👉📞 Just📲 Call Ruhi Colle...
👉 Chennai Sexy Aunty’s WhatsApp Number 👉📞 7427069034 👉📞 Just📲 Call Ruhi Colle...
rajnisinghkjn
 
Call Girl in Chennai | Whatsapp No 📞 7427069034 📞 VIP Escorts Service Availab...
Call Girl in Chennai | Whatsapp No 📞 7427069034 📞 VIP Escorts Service Availab...Call Girl in Chennai | Whatsapp No 📞 7427069034 📞 VIP Escorts Service Availab...
Call Girl in Chennai | Whatsapp No 📞 7427069034 📞 VIP Escorts Service Availab...
amritaverma53
 
Guntur Call Girl Service 📞6297126446📞Just Call Divya📲 Call Girl In Guntur No ...
Guntur Call Girl Service 📞6297126446📞Just Call Divya📲 Call Girl In Guntur No ...Guntur Call Girl Service 📞6297126446📞Just Call Divya📲 Call Girl In Guntur No ...
Guntur Call Girl Service 📞6297126446📞Just Call Divya📲 Call Girl In Guntur No ...
Call Girls in Nagpur High Profile Call Girls
 
Cara Menggugurkan Kandungan Dengan Cepat Selesai Dalam 24 Jam Secara Alami Bu...
Cara Menggugurkan Kandungan Dengan Cepat Selesai Dalam 24 Jam Secara Alami Bu...Cara Menggugurkan Kandungan Dengan Cepat Selesai Dalam 24 Jam Secara Alami Bu...
Cara Menggugurkan Kandungan Dengan Cepat Selesai Dalam 24 Jam Secara Alami Bu...
Cara Menggugurkan Kandungan 087776558899
 
Russian Call Girls In Pune 👉 Just CALL ME: 9352988975 ✅❤️💯low cost unlimited ...
Russian Call Girls In Pune 👉 Just CALL ME: 9352988975 ✅❤️💯low cost unlimited ...Russian Call Girls In Pune 👉 Just CALL ME: 9352988975 ✅❤️💯low cost unlimited ...
Russian Call Girls In Pune 👉 Just CALL ME: 9352988975 ✅❤️💯low cost unlimited ...
chanderprakash5506
 
❤️ Chandigarh Call Girls☎️98151-579OO☎️ Call Girl service in Chandigarh ☎️ Ch...
❤️ Chandigarh Call Girls☎️98151-579OO☎️ Call Girl service in Chandigarh ☎️ Ch...❤️ Chandigarh Call Girls☎️98151-579OO☎️ Call Girl service in Chandigarh ☎️ Ch...
❤️ Chandigarh Call Girls☎️98151-579OO☎️ Call Girl service in Chandigarh ☎️ Ch...
Rashmi Entertainment
 

Recently uploaded (20)

Cardiac Output, Venous Return, and Their Regulation
Cardiac Output, Venous Return, and Their RegulationCardiac Output, Venous Return, and Their Regulation
Cardiac Output, Venous Return, and Their Regulation
 
Bhawanipatna Call Girls 📞9332606886 Call Girls in Bhawanipatna Escorts servic...
Bhawanipatna Call Girls 📞9332606886 Call Girls in Bhawanipatna Escorts servic...Bhawanipatna Call Girls 📞9332606886 Call Girls in Bhawanipatna Escorts servic...
Bhawanipatna Call Girls 📞9332606886 Call Girls in Bhawanipatna Escorts servic...
 
👉 Chennai Sexy Aunty’s WhatsApp Number 👉📞 7427069034 👉📞 Just📲 Call Ruhi Colle...
👉 Chennai Sexy Aunty’s WhatsApp Number 👉📞 7427069034 👉📞 Just📲 Call Ruhi Colle...👉 Chennai Sexy Aunty’s WhatsApp Number 👉📞 7427069034 👉📞 Just📲 Call Ruhi Colle...
👉 Chennai Sexy Aunty’s WhatsApp Number 👉📞 7427069034 👉📞 Just📲 Call Ruhi Colle...
 
(RIYA)🎄Airhostess Call Girl Jaipur Call Now 8445551418 Premium Collection Of ...
(RIYA)🎄Airhostess Call Girl Jaipur Call Now 8445551418 Premium Collection Of ...(RIYA)🎄Airhostess Call Girl Jaipur Call Now 8445551418 Premium Collection Of ...
(RIYA)🎄Airhostess Call Girl Jaipur Call Now 8445551418 Premium Collection Of ...
 
Circulatory Shock, types and stages, compensatory mechanisms
Circulatory Shock, types and stages, compensatory mechanismsCirculatory Shock, types and stages, compensatory mechanisms
Circulatory Shock, types and stages, compensatory mechanisms
 
💰Call Girl In Bangalore☎️63788-78445💰 Call Girl service in Bangalore☎️Bangalo...
💰Call Girl In Bangalore☎️63788-78445💰 Call Girl service in Bangalore☎️Bangalo...💰Call Girl In Bangalore☎️63788-78445💰 Call Girl service in Bangalore☎️Bangalo...
💰Call Girl In Bangalore☎️63788-78445💰 Call Girl service in Bangalore☎️Bangalo...
 
Call Girl in Chennai | Whatsapp No 📞 7427069034 📞 VIP Escorts Service Availab...
Call Girl in Chennai | Whatsapp No 📞 7427069034 📞 VIP Escorts Service Availab...Call Girl in Chennai | Whatsapp No 📞 7427069034 📞 VIP Escorts Service Availab...
Call Girl in Chennai | Whatsapp No 📞 7427069034 📞 VIP Escorts Service Availab...
 
Call 8250092165 Patna Call Girls ₹4.5k Cash Payment With Room Delivery
Call 8250092165 Patna Call Girls ₹4.5k Cash Payment With Room DeliveryCall 8250092165 Patna Call Girls ₹4.5k Cash Payment With Room Delivery
Call 8250092165 Patna Call Girls ₹4.5k Cash Payment With Room Delivery
 
Call Girls Bangalore - 450+ Call Girl Cash Payment 💯Call Us 🔝 6378878445 🔝 💃 ...
Call Girls Bangalore - 450+ Call Girl Cash Payment 💯Call Us 🔝 6378878445 🔝 💃 ...Call Girls Bangalore - 450+ Call Girl Cash Payment 💯Call Us 🔝 6378878445 🔝 💃 ...
Call Girls Bangalore - 450+ Call Girl Cash Payment 💯Call Us 🔝 6378878445 🔝 💃 ...
 
Bhopal❤CALL GIRL 9352988975 ❤CALL GIRLS IN Bhopal ESCORT SERVICE
Bhopal❤CALL GIRL 9352988975 ❤CALL GIRLS IN Bhopal ESCORT SERVICEBhopal❤CALL GIRL 9352988975 ❤CALL GIRLS IN Bhopal ESCORT SERVICE
Bhopal❤CALL GIRL 9352988975 ❤CALL GIRLS IN Bhopal ESCORT SERVICE
 
Call Girls Rishikesh Just Call 9667172968 Top Class Call Girl Service Available
Call Girls Rishikesh Just Call 9667172968 Top Class Call Girl Service AvailableCall Girls Rishikesh Just Call 9667172968 Top Class Call Girl Service Available
Call Girls Rishikesh Just Call 9667172968 Top Class Call Girl Service Available
 
Indore Call Girls ❤️🍑7718850664❤️🍑 Call Girl service in Indore ☎️ Indore Call...
Indore Call Girls ❤️🍑7718850664❤️🍑 Call Girl service in Indore ☎️ Indore Call...Indore Call Girls ❤️🍑7718850664❤️🍑 Call Girl service in Indore ☎️ Indore Call...
Indore Call Girls ❤️🍑7718850664❤️🍑 Call Girl service in Indore ☎️ Indore Call...
 
Guntur Call Girl Service 📞6297126446📞Just Call Divya📲 Call Girl In Guntur No ...
Guntur Call Girl Service 📞6297126446📞Just Call Divya📲 Call Girl In Guntur No ...Guntur Call Girl Service 📞6297126446📞Just Call Divya📲 Call Girl In Guntur No ...
Guntur Call Girl Service 📞6297126446📞Just Call Divya📲 Call Girl In Guntur No ...
 
Cara Menggugurkan Kandungan Dengan Cepat Selesai Dalam 24 Jam Secara Alami Bu...
Cara Menggugurkan Kandungan Dengan Cepat Selesai Dalam 24 Jam Secara Alami Bu...Cara Menggugurkan Kandungan Dengan Cepat Selesai Dalam 24 Jam Secara Alami Bu...
Cara Menggugurkan Kandungan Dengan Cepat Selesai Dalam 24 Jam Secara Alami Bu...
 
Chennai Call Girls Service {7857862533 } ❤️VVIP ROCKY Call Girl in Chennai
Chennai Call Girls Service {7857862533 } ❤️VVIP ROCKY Call Girl in ChennaiChennai Call Girls Service {7857862533 } ❤️VVIP ROCKY Call Girl in Chennai
Chennai Call Girls Service {7857862533 } ❤️VVIP ROCKY Call Girl in Chennai
 
ANATOMY AND PHYSIOLOGY OF REPRODUCTIVE SYSTEM.pptx
ANATOMY AND PHYSIOLOGY OF REPRODUCTIVE SYSTEM.pptxANATOMY AND PHYSIOLOGY OF REPRODUCTIVE SYSTEM.pptx
ANATOMY AND PHYSIOLOGY OF REPRODUCTIVE SYSTEM.pptx
 
Call Girls Service Jaipur {9521753030 } ❤️VVIP BHAWNA Call Girl in Jaipur Raj...
Call Girls Service Jaipur {9521753030 } ❤️VVIP BHAWNA Call Girl in Jaipur Raj...Call Girls Service Jaipur {9521753030 } ❤️VVIP BHAWNA Call Girl in Jaipur Raj...
Call Girls Service Jaipur {9521753030 } ❤️VVIP BHAWNA Call Girl in Jaipur Raj...
 
Russian Call Girls In Pune 👉 Just CALL ME: 9352988975 ✅❤️💯low cost unlimited ...
Russian Call Girls In Pune 👉 Just CALL ME: 9352988975 ✅❤️💯low cost unlimited ...Russian Call Girls In Pune 👉 Just CALL ME: 9352988975 ✅❤️💯low cost unlimited ...
Russian Call Girls In Pune 👉 Just CALL ME: 9352988975 ✅❤️💯low cost unlimited ...
 
💞 Safe And Secure Call Girls Coimbatore🧿 6378878445 🧿 High Class Coimbatore C...
💞 Safe And Secure Call Girls Coimbatore🧿 6378878445 🧿 High Class Coimbatore C...💞 Safe And Secure Call Girls Coimbatore🧿 6378878445 🧿 High Class Coimbatore C...
💞 Safe And Secure Call Girls Coimbatore🧿 6378878445 🧿 High Class Coimbatore C...
 
❤️ Chandigarh Call Girls☎️98151-579OO☎️ Call Girl service in Chandigarh ☎️ Ch...
❤️ Chandigarh Call Girls☎️98151-579OO☎️ Call Girl service in Chandigarh ☎️ Ch...❤️ Chandigarh Call Girls☎️98151-579OO☎️ Call Girl service in Chandigarh ☎️ Ch...
❤️ Chandigarh Call Girls☎️98151-579OO☎️ Call Girl service in Chandigarh ☎️ Ch...
 

British J Pharmacology - 2021 - Healy - The importance of Ras in drug resistance in cancer.pdf

  • 1. Themed Issue: New avenues in cancer prevention and treatment (BJP 75th Anniversary) C O M M I S S I O N E D R E V I E W A R T I C L E - T H E M E D I S S U E The importance of Ras in drug resistance in cancer Fiona M. Healy1 | Ian A. Prior2 | David J. MacEwan1 1 Department of Pharmacology and Therapeutics, Institute of Systems, Molecular and Integrative Biology (ISMIB), University of Liverpool, Liverpool, UK 2 Department of Molecular Physiology and Cell Signalling, Institute of Systems, Molecular and Integrative Biology (ISMIB), University of Liverpool, Liverpool, UK Correspondence Prof. David J. MacEwan, Department of Pharmacology and Therapeutics, Institute of Systems, Molecular and Integrative Biology (ISMIB), University of Liverpool, Liverpool L69 3GE, UK. Email: macewan@liverpool.ac.uk In this review, we analyse the impact of oncogenic Ras mutations in mediating cancer drug resistance and the progress made in the abrogation of this resistance, through pharmacological targeting. At a physiological level, Ras is implicated in many cellular proliferation and survival pathways. However, mutations within this small GTPase can be responsible for the initiation of cancer, therapeutic resistance and failure, and ultimately disease relapse. Often termed “undruggable,” Ras is notoriously difficult to target directly, due to its structure and intrinsic activity. Thus, Ras-mediated drug resistance remains a considerable pharmacological problem. However, with advances in both analytical techniques and novel drug classes, the therapeutic landscape against Ras is changing. Allele-specific, direct Ras-targeting agents have reached clinical trials for the first time, indicating there may, at last, be hope of targeting such an elusive but significant protein for better more effective cancer therapy. LINKED ARTICLES: This article is part of a themed issue on New avenues in cancer prevention and treatment (BJP 75th Anniversary). To view the other articles in this section visit http://onlinelibrary.wiley.com/doi/10.1111/bph.v179.12/issuetoc K E Y W O R D S acquired resistance, cancer stem cells, Drug resistance, intrinsic resistance, MAPK, PI3K 1 | INTRODUCTION Resistance to conventional therapeutic agents is an increasingly con- cerning issue across all areas of disease, including cancer. Whilst the heterogeneous nature of cancer means there are many mechanisms resulting in drug resistance, Ras mutations underpin resistance to a variety of therapies (Hobbs et al., 2016; Prior et al., 2012, 2020). Oncogenic mutations in this small GTPase, which occur in approxi- mately 19% of all cancers, cause constitutive activation of prolifera- tive and survival pathways (Prior et al., 2020). This can abrogate the effects of standard chemotherapy and newer receptor-targeted therapies. Examples of such resistance is seen across a wide range of cancers, including metastatic colorectal cancer (mCRC), non-small cell lung cancer (NSCLC), pancreatic cancer, acute myeloid leukaemia (AML) and basal cell carcinoma (Li et al., 2004; McMahon et al., 2019; Misale et al., 2012; Tao et al., 2014). Thus, there is a distinct clinical need to target Ras pharmacologically. Whilst this has been particularly challenging due to structural difficulties and very high levels of intrin- sic activity, significant developments have been made within the last decade. Allele-specific, direct Ras-targeting reaching clinical trials pre- sent a new era of potential Ras therapeutics and increase the likeli- hood of overcoming this resistance mechanism. Abbreviations: EGFR, epidermal growth factor receptor; FLT3, fms related receptor tyrosine kinase 3; GAP, GTPase-activating protein; GEF, guanine nucleotide exchange factor; lncRNAs, long non-coding RNAs; MEK, MAPK kinase; RTK, receptor tyrosine kinase; SOS, son-of-sevenless; SOS1, SOS Ras/Rac guanine nucleotide exchange factor 1; TKI, tyrosine kinase inhibitor. Received: 17 December 2020 Revised: 10 February 2021 Accepted: 21 February 2021 DOI: 10.1111/bph.15420 This is an open access article under the terms of the Creative Commons Attribution License, which permits use, distribution and reproduction in any medium, provided the original work is properly cited. © 2021 The Authors. British Journal of Pharmacology published by John Wiley & Sons Ltd on behalf of British Pharmacological Society. 2844 Br J Pharmacol. 2022;179:2844–2867. wileyonlinelibrary.com/journal/bph
  • 2. 2 | WHAT IS DRUG RESISTANCE? There are two general classes of resistance, intrinsic and acquired. Intrinsic resistance occurs due to overt pre-existing factors, including variations in protein expression levels (such as increased expression of the P-gp [MDR1] transporter), epigenetic modifications (including by the long non-coding RNA (lncRNA) HAND2-AS1 and chromatin modi- fier Jarid1A (lysine demethylase 5A)) and somatic mutations (such as in Ras) (Burrell et al., 2013; Gruber et al., 2012; Marusyk et al., 2012; Sharma et al., 2010). Changes conferring drug resistance often co- exist, resulting in a resistance heterogeneity similar to that seen in the original disease itself (Gerlinger et al., 2012; Ramirez et al., 2016). Acquired resistance is thought to occur for a number of reasons, including through pre-existing (but initially undetectable) and de novo mutations (Bhaduri et al., 2020; Russo et al., 2019). This is often identified weeks to months after treatment has commenced (Santoni- Rugiu et al., 2019). In recent times, the concept of disease clonal heterogeneity has provided greater insight into the causes of acquired resistance, suggesting that chemoresistance and disease relapse occur as a result of minor subclonal populations. Given that these likely contribute to the heterogeneous nature of cancer, it seems likely that these also contribute to disease re-emergence and relapse (Bonnet & Dick, 1997; Gerlinger et al., 2012; Pattabiraman & Weinberg, 2014; Roy & Cowden Dahl, 2018; Seth et al., 2019). Within these minor clonal populations, mutations conferring drug resistance may exist at the early stages of disease but remain dormant and undetectable upon first presentation (Pietrantonio et al., 2017; Russo et al., 2018). When the bulk of the cancer is eliminated as a result of initial chemotherapy targeted at overt mutations, cells from this minor subclone proliferate and become dominant in the tumour bulk (Jones et al., 2019; McMahon et al., 2019). A second, related, resistance mechanism involves a very rare subpopulation of cells, known as cancer stem cells. Cancer stem cells were first described in acute myeloid leukaemia but have since been applied to many other cancers including (but not limited to) breast cancer, colorectal cancer, myeloma and pancreatic adenocarcinoma (Bonnet & Dick, 1997; Koury et al., 2017; Lapidot et al., 1994). Cancer stem cells have unique properties compared to bulk tumour cells. They are undifferentiated and have strong self-renewal and proliferative capabilities and typically remain in a quiescent state, thus avoiding chemotherapy targeted at rapidly dividing cells (as in bulk tumour cells). However, they do have the proliferative capacity to maintain and expand the tumour burden, as bulk tumour cells are eliminated (Jordan et al., 2006). These “stemness characteristics” render this subset of cells intrinsically resistant to chemotherapy, with distinct immunophenotypic and molecular signatures (Bonnet & Dick, 1997). This includes increased expression of efflux trans- porter P-gp (MDR1) and the ability to repair damaged DNA, a common method of inducing cancer cell death (Dean et al., 2005; Pattabiraman & Weinberg, 2014). Certain pathways up-regulated in cancer stem cells have been linked to the increased self-renewal capacity seen in this subset of cells. Key examples include the Wnt/β-catenin and Hedgehog (Hh) pathways, as well as the Ras-dependent MAPK and PI3K/AKT pathways. Of these, the Wnt/β-catenin pathway is perhaps the most associated with cancer stem cells. Briefly, at a physiological level, Wnt signalling is not active and β-catenin is ubiquitinated and sent for proteasomal degradation following interaction with GSK3β, APC and Axin. In cancer stem cells however, activation of Wnt signalling inhibits formation of the β-catenin–GSK3β–APC–Axin complex, stabilising β-catenin. This translocates to the nucleus whereby it stimulates transcription of proliferative genes, including c-Myc. This promotes proliferative signalling, increasing the self-renewal capacity of cancer stem cells (Krausova & Korinek, 2014; Moon et al., 2014). Although still ambiguous, it has been reported that β-catenin stabilisation can promote Ras stabilisation and protects Ras against proteasomal degradation (Jeong et al., 2018; Lee et al., 2018). This in turn promotes MAPK and PI3K pathway signalling, further contribut- ing to the self-renewal capacity of cancer stem cells. Alternatively, up-regulation of the hedgehog (Hh) signalling pathway has been shown in cancer stem cells. Whilst there are many facets to this pathway, its activation can stimulate up-regulation of stemness-associated transcription factors including NANOG, OCT4 and SOX2, thus promoting self-renewal (Boyer et al., 2005; Po et al., 2010; Takahashi & Yamanaka, 2006; Zhu et al., 2019). Increased activation and stabilisation of these genes (through either Hh signal- ling or biochemical stresses) have been implicated in the dedifferentia- tion of bulk tumour cells to cancer stem cells (Herreros-Villanueva et al., 2013; Kumar et al., 2012). Interest in epigenetic remodelling has grown considerably in the last decade and it has considerable effects in promoting drug resis- tance. lncRNAs are also implicated in survival of cancer stem cells. In hepatocellular carcinoma, the lncRNA HAND2-AS1 is up-regulated and stimulates the self-renewal capacity of cancer stem cells, through activation of the bone morphogenetic protein (BMP) signalling path- way. BMP activation has in turn been shown to induce chemoresistance in other cancer models, such as lung cancer (Gruber et al., 2012; Wang et al., 2015). Indeed, this pathway is also regulated by the fusion gene CBFA2T3-GLIS2 and the Hh and JAK-STAT signal- ling pathways (Gruber et al., 2012; Okada et al., 2014). Taken together, it is evident that many signalling pathways and associated factors play a critical role in mediating drug resistance. Identifying commonalities between these pathways could present an opportunity to reduce the incidence and impact of drug resistance. Although the incidence of Ras mutations in cancer stem cells is rela- tively low, Ras-mediated pathways have been implicated (Corces- Zimmerman et al., 2014). In addition to the aforementioned β-catenin- mediated stabilisation of Ras, stabilised OCT4 and SOX2 expression has also been shown in response to AKT activation, by extracellular biochemical and radiation stresses (Maiuthed et al., 2018; Park et al., 2021). Indeed, activation of the MAPK pathway through the Ras–Raf interaction promotes increased expression of SOX2 and NANOG (Chan et al., 2018; Du et al., 2019). Furthermore, correlation has been seen between up-regulation of the MAPK, PI3K and BMP pathways in drug-resistant lung cancer (Wang et al., 2015). Thus, not only can these pathways contribute to an increased cellular HEALY ET AL. 2845 14765381, 2022, 12, Downloaded from https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420, Wiley Online Library on [11/06/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
  • 3. proliferation rate in bulk cancer, but they can also assist with the self- renewal and stemness properties evident in cancer stem cells. Perhaps unsurprisingly, inhibition of the Hh signalling pathway, when com- bined with inhibition of mTOR (downstream of AKT), has been seen to eliminate cancer stem cells in pancreatic cancer (Mueller et al., 2009; Prieur et al., 2017). Furthermore, MAPK kinase (MEK) inhibition has been implicated in reduced pancreatic cancer stem cell survival (Walter et al., 2019). 3 | KEY EXAMPLES OF DRUG RESISTANCE IN CANCER In recent years, resistance has been documented amongst many can- cer therapeutics, including traditional chemotherapeutic agents, more novel small molecule inhibitors and monoclonal antibodies (Caiola et al., 2015; McMahon et al., 2019; Pietrantonio et al., 2017). The origin of this resistance (intrinsic or acquired) varies considerably between drugs and so advances have been made in treatment stratifi- cation, based on a patient's mutational status, for some therapies. For example, vemurafenib, the BRAF inhibitor, is restricted to melanoma patients with the BRAF V600E mutation (Hopkins et al., 2019). Like- wise, the monoclonal antibody panitumumab is only recommended for patients with wild-type Ras (Amado et al., 2008). However, as mentioned previously, some mutations are only detectable when patients relapse, such as the emergence of fms-like TK 3 (FLT3) or NRAS mutations in acute myeloid leukaemia patients (Man et al., 2012; McMahon et al., 2019; Piloto et al., 2007; Smith et al., 2017). Many of these mutations occur downstream of the site targeted by the drug. For example, since Ras operates downstream of the EGF receptor (EGFR), Ras mutations play a role in rendering cetuximab and panitumumab (which bind and inhibit EGFR) ineffec- tive as the constitutive activation is not inhibited by the drug (De Roock et al., 2010; Li et al., 2015; Zhao et al., 2017). However, some of these putative resistance-causing mutations can only be detected at relapse, once they have expanded from only existing in a minor, undetectable subclone (as they did at diagnosis). Thus, it is difficult to predict which patients will develop these resistance mutations, meaning initial treatment stratification is difficult. Thus, combination therapy between these established agents and novel agents targeting these common mutational sites may be a way forward in preventing resistance before it occurs, an example of which is the combination of BRAF and MEK inhibitors, vemurafenib and cobimetinib (Hopkins et al., 2019). 4 | INTRODUCTION TO Ras As eluded, Ras mutations underpin resistance to a variety of therapies. KRAS is most commonly mutated of the three Ras isoforms and is particularly frequently mutated in lung and pancreatic cancer (Moore, Rosenberg, et al., 2020; Prior et al., 2020). However, patterns of Ras mutation differ between cancers and NRAS is the most frequently mutated RAS gene in acute myeloid leukaemia. Also, HRAS accounts for most Ras mutations in head and neck squamous cell carcinoma (Prior et al., 2020). Indeed, this variation carries varying prognoses of Ras-mutated cancer, with KRAS mutations conferring the lowest over- all survival 5 years post diagnosis (43%) and HRAS mutations confer- ring the highest (63%) (Figure 1) (Cerami et al., 2012; Gao et al., 2013). Most oncogenic mutations in Ras occur in the residues G12, G13 and Q61, which are located in a region conserved between HRAS, KRAS and NRAS (Figure 2a). They occur within the effector lobe, which comprises the first 85 amino acids and is a region of complete sequence identity between the different RAS isoforms. Most key interaction sites occur within the effector lobe, including nucleotide and effector interaction sites, as well as the switch regions that medi- ate effector interactions (Figure 2b). There is 90% similarity in the next 80 amino acids (allosteric lobe). The only considerable sequence differences occur at the C terminal end of the protein, known as the hypervariable region. This is the region in which post-translational modifications occur and membrane-targeting sequences are found (Hobbs et al., 2016). Initial attempts at direct pharmacological targeting of Ras centred around inhibiting these post-translational modifications, including farnesylation (Cox & Der, 2002; Whyte et al., 1997). However, more recent attempts have considered the effector lobe and structures within this to be better therapeutic targets, as discussed later (Canon et al., 2019; Janes et al., 2018; Ostrem et al., 2013). Ras is activated or inactivated when bound to GTP or GDP, respectively. Oncogenic mutations increase the frequency of GTP- bound (active) Ras, through two key mechanisms. This is either through decreasing the affinity of Ras for GTPase-activating proteins (GAPs), which stimulate the intrinsic GTPase activity of Ras and there- fore facilitate hydrolysis of GTP to GDP (off/inactivating mechanism) or by decreasing the need for guanine nucleotide exchange factors (GEFs), which mediate the on/activating mechanism (Smith et al., 2013; Wittinghofer & Waldmann, 2000). These differences depend on multiple factors including the specific amino acid mutation and resulting protein conformation (Hunter et al., 2015; Miller & Miller, 2012; Poulin et al., 2019; Prior et al., 2012; Smith et al., 2013). Some mutations (including G12C) promote rapid cycling between the inactive and active states, an area which has been exploited by a range of novel Ras-targeting drugs (Fell et al., 2020; Hunter et al., 2015; Patricelli et al., 2016). 5 | Ras SIGNALLING—REGULAR AND ONCOGENIC Ras is involved in key pathways regulating cell proliferation, differentiation and sensitivity to apoptosis, perhaps the most notable being the MAPK pathway and the PI3K/AKT pathway. KRAS has also been implicated (albeit in a more indirect way) in a range of other pathways, including the Wnt/β-catenin pathway and the nuclear factor erythroid 2-related factor 2 (NRF2) pathway (DeNicola 2846 HEALY ET AL. 14765381, 2022, 12, Downloaded from https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420, Wiley Online Library on [11/06/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
  • 4. et al., 2011; Ferino et al., 2020; Moon et al., 2014; Park et al., 2019; Tao et al., 2014). These pathways are detailed briefly below and are summarised in Figure 3. Ligand binding to receptor tyrosine kininases (RTKs), such as EGFR or FLT3, causes autophosphorylation of key tyrosine residues on the RTK, after which adaptor proteins such as GRB2 bind to these phosphorylated tyrosines via SH2 domains. Son-of-sevenless (SOS), a Ras-guanine nucleotide exchange factor, then associates with GRB2 via two SH3 domains and ultimately facilitates the exchange of GDP for GTP on RAS, thereby activating it (Freeman, 2000). Following this, Ras stimulates recruitment of its serine–threonine kinase effector, RAF, to the membrane, which binds through its Ras binding domain and then proceeds to phosphorylate downstream kinases including MEK and ERK (Marais et al., 1995; Molina & Adjei, 2006). ERK then translocates from the cytoplasm to the nucleus, causing the activation of many transcription factors, which FIGURE 1 The impact of Ras alterations on disease. Data collected from 32 curated, non-redundant studies, comprising 10,967 patient samples, as collated by the TCGA Pan-Cancer Atlas. (a) Ten-year profession-free survival (PFS) analysis for patients with and without Ras alterations. Ras wild-type (blue) versus altered (red) overall survival. Ras wild-type median PFS = 65.88 months, Ras-altered median PFS = 44.19 months. P < .01. (b) Ten-year PFS stratified by altered Ras isoform. KRAS (orange) median PFS = 36.72 months, NRAS (green) median PFS = 45.67 months, HRAS (purple) median PFS = 74.89 months. P < .01. (c) Cancer types where KRAS is altered in at least 10% of cases. (d) Cancer types where NRAS is altered in at least 10% of cases. (e) Cancer types where HRAS is altered in at least 10% of cases. Data obtained from cBioPortal, all TCGA Pan-Cancer Atlas Studies HEALY ET AL. 2847 14765381, 2022, 12, Downloaded from https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420, Wiley Online Library on [11/06/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
  • 5. T A B L E 1 Summary of current Ras pathway targeting drugs, at varying stages of development Name Drug type Target Development stage Mode of inhibition Selected common adverse effects References Gilteritinib Pyrazinecarboxamide, small molecule inhibitor FLT3-ITD, FLT3-TKD, AXL FDA-approved Binds active FLT3 (either ITD or TKD), inhibiting constitutive signalling Acute kidney injury, hypotension, diarrhoea, dizziness Lee et al. (2017); Perl et al. (2019) Midostaurin Indolocarbazole, small molecule inhibitor FLT3 FDA-approved Binds active FLT3, inhibiting constitutive signalling Nausea, neutropenia, thrombocytopenia, diarrhoea, vomiting Levis (2017); Stone et al. (2012) Crenolanib Benzamidazole-derivative small molecule inhibitor FLT3-ITD, FLT3-TKD Phase II (NCT02400255) Binds active FLT3, inhibiting constitutive signalling Nausea, vomiting, diarrhoea, stomach pain Galanis et al. (2014); Marensi et al. (2021); Wu et al. (2018) Cetuximab Chimeric IgG1 chimeric monoclonal antibody EGFR FDA-approved Competitively binds extracellular domain of EGFR, preventing ligand-mediated signalling Oedema, fatigue, anorexia, rash, vomiting Jonker et al. (2007) Panitumumab Human IgG2 monoclonal antibody EGFR FDA-approved Competitively binds extracellular domain of EGFR, inhibiting EGFR dimerization and autophosphorylation Rash, diarrhoea, hypomagnesaemia Van Cutsem et al. (2007) BAY293 Quinazoline-based small molecule inhibitor SOS1 Preclinical Disruption of the KRAS–SOS1 interaction, reducing exchange of GDP for GTP N/A Hillig et al. (2019) SAH-SOS1 Stapled peptide fragment SOS1 Preclinical Binds KRAS (active or inactive) in the SOS1 binding pocket, reducing GTP binding to KRAS N/A Leshchiner et al. (2015) BI1701963 Small molecule inhibitor SOS1 Phase I (NCT04111458) Binds inside the catalytic site of SOS1, preventing interaction with (and activation of) KRAS N/A Gerlach et al. (2020) KYA1797K Thiazolidine-based small molecule inhibitor Axin Preclinical Binds to the RGS domain of Axin, causing formation of the β-catenin destruction complex, subsequently destabilising Ras N/A Cha et al. (2016); Lee et al. (2018) Tipifarnib Quinolinone small molecule inhibitor Farnesyltransferase FDA-approved Prevents Ras farnesylation and subsequent trafficking to the membrane Thrombocytopenia, anorexia, anaemia, nausea, neutropenia Duffy and Crown (2021); Lee et al. (2020) ARS-853 Acrylamide-based, small molecule inhibitor KRAS G12C Preclinical Irreversible binding to KRAS G12C SII-P pocket, inhibiting exchange of GDP for GTP N/A Lito et al. (2016); Patricelli et al. (2016) 2848 HEALY ET AL. 14765381, 2022, 12, Downloaded from https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420, Wiley Online Library on [11/06/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
  • 6. T A B L E 1 (Continued) Name Drug type Target Development stage Mode of inhibition Selected common adverse effects References ARS-1620 Acrylamide-based, small molecule inhibitor (S-atropisomer) KRAS G12C Preclinical Irreversible binding to KRAS G12C SII-P pocket, inhibiting exchange of GDP for GTP N/A Janes et al. (2018) AMG-510 Acrylamide-based, small molecule inhibitor KRAS G12C Phase I/II (NCT03600883) Irreversible binding to KRAS G12C SII-P pocket, inhibiting exchange of GDP for GTP Anaemia, diarrhoea Canon et al. (2019); Govindan (2019) MRTX849 Acrylamide-based, small molecule inhibitor KRAS G12C Phase I (NCT03785249) Covalent binding in SII-P pocket, inhibiting exchange of GDP for GTP Nausea, diarrhoea, fatigue, hyponatremia Fell et al. (2020); Hallin et al. (2020a) inRas37 Human IgG1 internalising and PPI-interfering monoclonal antibody pan-Ras Preclinical Competitive inhibition of RAS– effector interaction N/A Shin et al. (2020) siG12D- LODER Long-acting siRNA KRAS G12D Phase II (NCT01676259) siRNA-mediated KRAS G12D silencing Diarrhoea, abdominal pain, nausea, fatigue Golan et al. (2015); Zorde Khvalevsky et al. (2013) Dabrafenib Sulphonamide-based small molecule inhibitor BRAF (wild-type and V600-mutated) FDA-approved ATP-competitive inhibitor, binds active BRAF and thus inhibits downstream effector activation Rash, fever, fatigue, headache, hypertension, arthralgia Bowyer et al. (2015); King et al. (2013); Rheault et al. (2013) Vemurafenib Azaindole-derived small molecule inhibitor BRAF V600E FDA-approved ATP-competitive inhibitor, binds active BRAF and thus inhibits downstream receptor activation Skin lesions (including squamous cell carcinoma), arthralgia, fatigue Sharma et al. (2012); Tsai et al. (2008) Sorafenib Biaryl-urea-based small molecule inhibitor RAF, PDGFR, VEGFR amongst others FDA-approved Binds inactive conformation of BRAF, reducing activation. Also sequesters Raf into inactive complexes Diarrhoea, weight loss, skin reactions, alopecia, voice changes Adnane et al. (2006); Llovet et al. (2008); Marensi et al. (2021); Wan et al. (2004); Wilhelm et al. (2006) Cobimetinib Carboxamide-based small molecule inhibitor RAF, MEK FDA-approved Non-ATP-competitive inhibitor, binds active MEK, inhibiting downstream ERK activation Diarrhoea, rash, fatigue, arthralgia, photosensitivity Garnock-Jones (2015); Hatzivassiliou et al. (2013); Rice et al. (2012) Trametinib Pyridopyrimidine small molecule inhibitor MEK FDA-approved ATP non-competitive kinase inhibitor, binds the kinase suppressor of RAS–RAS interface, reducing MEK phosphorylation Diarrhoea, rash, blurred vision Borthakur et al. (2016); Khan et al. (2020) LY3214996 Thiazolone-based small molecule inhibitor ERK Phase I (NCT02857270) ATP-competitive inhibitor, reversibly binds ERK1 and ERK2, causes cell cycle arrest in G 1 and initiates apoptosis Nausea, vomiting, diarrhoea, fatigue, blurred vision Bhagwat et al. (2020); Pant et al. (2019) (Continues) HEALY ET AL. 2849 14765381, 2022, 12, Downloaded from https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420, Wiley Online Library on [11/06/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
  • 7. go on to regulate a host of processes, including proliferation and dif- ferentiation (Guo et al., 2020; Zhang & Liu, 2002). Alternatively, activated Ras can promote the PI3K–AKT pathway. This pathway also plays a key role in controlling survival, division and metabolism and can be activated in many different ways (Castellano & Downward, 2011; Mendoza et al., 2011). Ras binds and activates the p110α catalytic subunit of PI3K, which in turn promotes transforma- tion of the plasma membrane-bound lipid PIP2 to PIP3 and activation of AKT by binding through its pleckstrin homology (PH) domain (Can- tley, 2002; Castellano & Downward, 2011). AKT interacts with many downstream effectors, including MDM2, BAX BAD and NF-κB. These interactions regulate apoptosis (Cantley, 2002; Castellano & Down- ward, 2011; Chang et al., 2003; Duronio, 2008). Furthermore, activa- tion of FOXO by AKT promotes cellular metabolism (Engelman et al., 2006). Given that the MAPK and PI3K pathways are both strongly involved in controlling cell survival and death, it is clear that aberrant signalling within these pathways will lead to cancer, of which key hallmarks include sustaining proliferative signalling, enabling repli- cative immortality and resisting cell death (Hanahan & Weinberg, 2011). Aside from these two classic Ras-dependent pathways, increasing evidence suggests the implication of normal and mutant Ras function in alternative mechanisms, such as redox homeostasis and stem cell survival (Mukhopadhyay et al., 2020; Wu et al., 2019). The role of Ras in cancer stem cells is yet to be fully elucidated. However, its crosstalk with the Wnt/β-catenin pathway in tumorigenesis may suggest its involvement in the maintenance of cancer stem cells (Moon et al., 2014). It is understood that, in colorectal cancer, there is over- activation of both the Wnt/β-catenin and Ras pathways, through mutations within APC (a tumour suppressor) and KRAS, respectively (Jeong et al., 2018). There is a synergistic effect seen with these mutations (D'Abaco et al., 1996; Janssen et al., 2006; Jeong et al., 2018; Margetis et al., 2017). Mutations within APC cause loss of function of the tumour suppressor gene, whilst KRAS mutations lead to phosphorylation of key tyrosine residues within β-catenin, causing its accumulation within the cytoplasm. This ultimately increases acti- vation of downstream Wnt pathway target genes, including REG4, a marker of cancer stem cells (Hwang et al., 2020; Janssen et al., 2006). Up-regulation of this pathway is highly associated with the increased survival and plasticity properties seen in cancer stem cells and has been seen in many cancers (Al-Hajj et al., 2003; Koury et al., 2017; Lapidot et al., 1994). Taken together, it seems plausible that these KRAS mutations may play a role in the protection of the minor subset of cells, which likely have a key role in relapse. Furthermore, Ras mutations can also be implicated in ROS generation/detoxification. ROS are considered a “double-edged sword,” capable of both helping and hindering cancer cells (Hayes & McMahon, 2006; Wu et al., 2019). At physiological levels of ROS, NRF2 binds Keap1 and is ubiquitinated and degraded on a regular basis. However, upon detection of high levels of ROS (which is carci- nogenic), NRF2 is unable to bind to Keap1 (due to conformational changes in Keap1) and so translocates to the nucleus, where it acts as a transcription factor for various downstream detoxification genes. T A B L E 1 (Continued) Name Drug type Target Development stage Mode of inhibition Selected common adverse effects References RBC8 Carbonitrile-based small molecule inhibitor RAL Preclinical Non-ATP-competitive inhibitor, binds GDP-loaded RAL, preventing activation N/A Walsh et al. (2019); Yan et al. (2014) Alpelisib Aminothiazole-based small molecule inhibitor PI3Kα FDA-approved ATP-competitive inhibitor, binds selectively to PI3Kα Hyperglycaemia, rash, diarrhoea, nausea, decreased appetite André et al. (2019); Furet et al. (2013) Uprosertib Thiophenecarboxemide-based small molecule inhibitor AKT Phase II (NCT01902173) ATP-competitive inhibitor, binds AKT and reduces downstream signalling Nausea, vomiting, diarrhoea, rash Gungor et al. (2015); Nitulescu et al. (2016) Everolimus Macrocyclic lactone-based small molecule inhibitor mTOR FDA-approved Complexes with FKBP12, which binds mTOR and inhibits activation Leukopenia, hypercholesterolaemia, hyperlipidaemia Dunn and Croom (2006) Abbreviations: PPI, protein–protein interaction; SOS, son-of-sevenless; 2850 HEALY ET AL. 14765381, 2022, 12, Downloaded from https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420, Wiley Online Library on [11/06/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
  • 8. FIGURE 2 2D and 3D representation of the structure of RAS. (a) Physiological binding domains. (b) Key structural domains (switch regions and lobes), with mutational hotspots G12, G13 and Q61 indicated. Redrawn from Prior et al. (2012). 3D structures based on PDB 4DST FIGURE 3 RAS-mediated pathways and associated inhibitors. Targets of small molecule inhibitors and monoclonal antibodies used across a range of cancers to inhibit proliferative signalling and survival of cancer cells. Figure includes examples of compounds identified in vitro, those which have progressed into trials and those which are approved. Further detail is provided in Table 1 HEALY ET AL. 2851 14765381, 2022, 12, Downloaded from https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420, Wiley Online Library on [11/06/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
  • 9. Hence, the DNA of the healthy cells remains undamaged by ROS (Basak et al., 2017; Gorrini et al., 2013). However, in cancer, the NRF2 pathway can be up-regulated to constitutively degrade ROS (induced by chemotherapeutics), thereby conferring a protective effect to the cancer cells, resulting in chemoresistance (Basak et al., 2017; Gorrini et al., 2013). Therefore, the balance of NRF2 activation level is crucial. In the last decade, it has been shown that KRAS G12D mutations can increase NRF2 transcription, through activation of the TRE (12-O- Tetradecanoylphorbol-13-acetate (TPA) response element) via the MAPK pathway (DeNicola et al., 2011; Mukhopadhyay et al., 2020; Shirazi et al., 2020; Tao et al., 2014). Thus, these KRAS mutations can render the cancer cell more capable of coping with chemotherapy- induced ROS, thereby mediating drug resistance. 6 | THE INVOLVEMENT OF Ras IN CHEMOTHERAPY RESISTANCE Should mutations occur as described above, it follows that one or many of these pathways can be perturbed, leading to increased cell proliferation, decreased cell death and promotion of cancer stem cells, amongst other effects. The following scenarios illustrate some of these key resistance mechanisms, highlighting the necessity for better Ras targeting. Platinum-based agents, such as cisplatin, carboplatin and oxaliplatin, are used in the treatment of a variety of cancers, including head and neck squamous cell carcinoma, testicular cancer and non-small cell lung cancer (de Vries et al., 2020; Silva et al., 2019; Weykamp et al., 2020). They are DNA intercalating agents that interfere with RNA transcription and DNA replication, through cross- linking of DNA. This results in the formation of DNA adducts, which in turn drive the tumour cell to apoptosis. Cisplatin also induces mitochondrial ROS, which further increase DNA damage and thus increase the cytotoxic properties of the drug (Marullo et al., 2013; Srinivas et al., 2019). However, there are many resistance mechanisms associated with cisplatin, including the involvement of oncogenic KRAS mutations (Caiola et al., 2015; DeNicola et al., 2011; Feldman et al., 2014; Garassino et al., 2011). KRAS mutations were shown to induce NRF2 pathway up-regulation in non-small cell lung cancer, thereby decreasing cisplatin-induced ROS within the tumour cell, and ultimately leading to decreased cell death (DeNicola et al., 2011). This was supported by further work indicating oncogenic KRAS can induce NRF2 gene transcription via the TPE response element, resulting in the overactivation of the anti-oxidative stress pathway, rendering the tumour cells resistant to cisplatin-induced ROS (Tao et al., 2014). Furthermore, KRAS mutations can lead to hyperactivation of the PI3K–AKT pathway, which is starting to be implicated as a cisplatin resistance mechanism. As mentioned, up-regulation of the PI3K– AKT–mTOR pathway can have multiple effects, including inhibition of apoptosis and increased cell proliferation (de Vries et al., 2020). Whilst there are other reasons for cisplatin resistance, Ras pathway mutations are heavily implicated in the key mechanisms. Thus, pharmacologically targeting Ras would provide an opportunity to overcome many causes of this resistance. Up-regulation of Ras-mediated pathways as a means of chemoresistance is by no means restricted to cisplatin resistance and is a common mechanism of resistance to TK inhibitors (TKIs), such as those targeting RTKs including FLT3 and EGFR (Eberlein et al., 2015; Massarelli et al., 2007; McMahon et al., 2019; Ortiz-Cuaran et al., 2016; Piloto et al., 2007; Van Emburgh et al., 2016). TKIs are used in a variety of cancers, including renal cell carcinoma (RCC), colorectal cancer, acute myeloid leukaemia and non- small cell lung cancer, to name a few examples. The mechanism of action of TKIs involves inhibition of phosphorylation sites within the protein, thereby preventing it exerting kinase activity on downstream effectors (Ciardiello & Tortora, 2008; Yamaoka et al., 2018). However, resistance to these can occur through two predominant mechanisms, mutations within the RTK or mutations within downstream pathways (McMahon et al., 2019; Piloto et al., 2007; Van Emburgh et al., 2016; Yu et al., 2013). Given that Ras occurs downstream of these recep- tors, any mutations within Ras will render the cell resistant to the TKI. For example, studies have shown KRAS mutations render patients resistant to gefitinib, used to treat non-small cell lung cancer (Pao et al., 2005; Zhao et al., 2017). In a similar way, the treatment of colorectal cancer with anti-EGFR monoclonal antibodies cetuximab or panitumumab is only successful in a subset of patients, with many eventually developing resistance (Pietrantonio et al., 2017). This has been attributed to Ras mutations and variations in the EGFR extracel- lular domain, which reduce antibody binding efficiency, ultimately initiating relapse (Van Emburgh et al., 2016). Although cetuximab and panitumumab are only prescribed to Ras wild-type patients, emer- gence of mutations from undetectable, pre-existing clones can give rise to resistance in this way, as evidenced through analysis of circu- lating tumour DNA (ctDNA) (Amirouchene-Angelozzi et al., 2017; Diaz et al., 2012; Misale et al., 2012). The order in which these mutations develop/emerge is likely important in understanding (and ultimately targeting) the process of relapse. Ras mutations often develop earlier than EGFR variations and typically confer poorer prognosis (Van Emburgh et al., 2016). Therefore, combatting these Ras mutations would not only improve prognosis of Ras-mutated patients but also provide a second therapeutic option for those that go on to develop extracellular domain variations. Resistance to FLT3-TKIs is also a highly prevalent issue. It is well documented that 20–30% of acute myeloid leukaemia patients have an internal tandem duplication in the FLT3 RTK (FLT3-ITD) causing increased cell proliferation and decreased apoptosis, via the MAPK, STAT5 and PI3K pathways (Hayakawa et al., 2000; Moore, Faisal, et al., 2020; Papaemmanuil et al., 2016; The Cancer Genome Atlas, 2013). Therefore, many different FLT3 inhibitors are at varying stages in development to overcome the effects of this mutation. Examples include gilteritinib, crenolanib and midostaurin (Aikawa et al., 2020; McMahon et al., 2019; Piloto et al., 2007; Zhang et al., 2019). These TKIs bind to the active conformation of FLT3 and are at varying stages of approval. Gilteritinib and midostaurin are FDA-approved while crenolanib is in Phase II trials (Galanis 2852 HEALY ET AL. 14765381, 2022, 12, Downloaded from https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420, Wiley Online Library on [11/06/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
  • 10. et al., 2014; Levis, 2017; Levis et al., 2011; Zhang et al., 2019). Although results for these drugs have all been promising, subsets of patients exhibit resistance. This has, in part, been attributed to Ras mutations, reactivating the MAPK and PI3K–AKT pathways (McMahon et al., 2019; Zhang et al., 2019). These mutations were detected in over 30% of patients who developed resistance to gilteritinib, with Ras variant allele frequencies also increasing post- drug exposure in patients who responded poorly to crenolanib. Inter- estingly, not all resistant patients had FLT3 mutations following treat- ment either, with different mutational signatures present instead (McMahon et al., 2019; Zhang et al., 2019). This implies a clonal selec- tion mechanism of resistance—a minor subpopulation at diagnosis which became dominant following treatment and elimination of the initial tumour burden. Taken together, it seems likely these Ras mutations, either pre-existing or de novo, may contribute to resistance to FLT3 inhibitors, through restoration of the original disease pheno- type by expansion of an originally minor subclone. This is perhaps unsurprising in acute myeloid leukaemia, which arises as a result of clonal haematopoiesis (Desai et al., 2018). 7 | OPTIONS FOR TARGETING THE Ras PATHWAY With improved capability of detecting minor cancer subclones, alongside the greater understanding of the impact of Ras mutations in various cancers, the next considerable challenge is improved pharma- cological targeting of Ras. However, drugging Ras has proven excep- tionally difficult. A key drawback is the lack of available binding sites for small molecule inhibitors. The nucleotide binding site (where GTP or GDP binds) seems a desirable pocket to target, however the picomolar affinity with which both GDP and GTP bind, as well as their high intracellular concentrations, effectively outcompetes the binding of any drug at this site (Cox et al., 2014; McCormick, 2018). Therefore, targeting alternative proteins within Ras-regulated pathways has been strongly investigated, with positive results seen, for example with trametinib, the MEK inhibitor. Currently approved for patients with BRAF V600-mutant metastatic melanoma or BRAF- mutated (V600) non-small cell lung cancer, trametinib is well tolerated in patients (Lugowska et al., 2015; Odogwu et al., 2018) and is also being assessed in Ras-mutant myeloid malignancies (Borthakur et al., 2016). This compound binds to phosphorylated MEK and inhibits its downstream effectors (e.g. ERK), despite the presence of constitutive Ras signalling. Subsequently, aberrant growth signalling and apoptosis inhibition is reduced (Hofmann et al., 2012). Whilst this has shown promising results (Borthakur et al., 2016; Lugowska et al., 2015; Odogwu et al., 2018), this compound only inhibits the MAPK pathway downstream of MEK, so constitutive activation of other Ras-dependent pathways (e.g. PI3K–AKT) will still occur in the presence of Ras mutations even when treated with this drug. In this way, cancer can persist (Jones et al., 2019; Stinchcombe & John- son, 2014). However, if Ras were to be targeted directly, signalling of both of these pathways would be inhibited, leading to cell death. As often seen with many diseases, combination of trametinib with other therapeutics to inhibit multiple pathways together may reduce the likelihood of continued cancer signalling and potential develop- ment of resistance to this and other drugs (Infante & Swanton, 2014; Planchard et al., 2016; Zhou et al., 2020). However, even with the approved combination regimen of dabrafenib (BRAF inhibitor) with trametinib (Lugowska et al., 2015), only the MAPK pathway is inhibited, thereby maintaining the potential for aberrant PI3K path- way signalling, which can in itself cause resistance to MEK inhibitors (Jaiswal et al., 2009; Sos et al., 2009; Vitiello et al., 2019). Alternatively, positive effects have been shown in vitro and in vivo of co-administering AKT inhibitors with dabrafenib to inhibit two Ras-mediated pathways (Lassen et al., 2014). However, combination of the two classes of drugs in patients did not yield significant clinical activity in reducing resistance seen in trametinib monotherapy, with 25% of patients exhibiting grade 3–4 toxicity (Algazi et al., 2018). Taken together, whilst downstream pathway inhibition has proved successful, this treatment method does have considerable disadvan- tages and may not be a long-term solution for many patients. Therefore, there is a distinct clinical need for novel means of treating Ras-mutant cancers, which could include the targeting of Ras itself. Given the difficulties with targeting Ras, inhibition of elements upstream of Ras has been investigated. This includes inhibition of GEFs including SOS Ras/Rac guanine nucleotide exchange factor 1 ( SOS1), to reduce the likelihood of Ras maintaining its GTP-bound state and therefore inhibiting constitutive signalling (Evelyn et al., 2014; Hillig et al., 2019). For example, BAY293 is a first-in-class compound with the ability to bind directly to SOS1 and inhibit the Ras–SOS interaction and thereby downstream signalling of the PI3K and MAPK pathways (Hillig et al., 2019). Although the in vivo bioavail- ability for this compound was poor, the concept of Ras–SOS inhibition could prove useful in the future, with promising high throughput in silico and in vitro screening results serving as a proof of concept for inhibition of Ras via this mechanism (Evelyn et al., 2014, 2015; Hillig et al., 2019). More recently, BI-1701963, a SOS1–pan-Ras interaction inhibitor, has reached Phase I clinical trials, the first of its kind to do so. Modified from the structure of BI-3406, a quinazoline-derived compound, this novel inhibitor binds to the catalytic site of SOS1, preventing its interaction with inactive KRAS, thus inhibiting activa- tion (Gerlach et al., 2020; Hofmann et al., 2020). In addition, alternative, more indirect pathways are also being targeted as a means of inhibiting Ras, which may also be able to elimi- nate the cancer stem cell. For example, the small molecule KYA1797K has been shown to be effective against oncogenic Ras in colorectal cancer and erlotinib-resistant non-small cell lung cancer (Park et al., 2019). This compound indirectly targets Ras, through inhibition of the Wnt/β-catenin pathway, which usually stabilises Ras (Jeong et al., 2012; Moon et al., 2014). This pathway is also up-regulated in cancer stem cells (Malanchi et al., 2008). KYA1797K has initiated anti- tumour effects in KRAS-mutant cell lines and a KRAS-mutated mouse model. KYA1797K also exhibited synergy with the current first-line therapeutic regimen (cisplatin and pemetrexed) in non-small cell lung cancer in vitro models (Park et al., 2019). However, KYA1797K HEALY ET AL. 2853 14765381, 2022, 12, Downloaded from https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420, Wiley Online Library on [11/06/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
  • 11. promoted apoptosis of both KRAS wild-type and mutant cells, questioning the specificity of the drug for cancer cells whilst sparing healthy cells. This study found both KRAS and β-catenin were overexpressed in tumour regions compared to non-tumour regions (Park et al., 2019). This may explain the limited toxicity seen during KYA1797K studies, with low doses of KYA1797K having a more sub- stantial effect on tumour cells, compared to healthy cells. Indeed, only limited toxicity was seen during in vivo studies (Park et al., 2019). However, protein expression varies considerably between tissues (with KRAS particularly highly expressed in the brain) and so targeting based on comparative expression between cancer and non-cancer regions in one tissue may not represent overall toxicity potential (Newlaczyl et al., 2017). Taken together, there may be a role for KYA1797K as a concomitant therapy in non-small cell lung cancer (as well as other cancers such as colorectal cancer). It presents a means of eliminating both the primary cause of the disease (if KRAS- mutated) and also minor subclones and cancer stem cells that could give rise to resistance (Cho et al., 2020). Nevertheless, better under- standing of the drug's effects on other tissues with high Ras expres- sion must be gained. 8 | OPTIONS FOR TARGETING Ras— DIRECT Ras TARGETING Ras post-translational modifications were targeted as a means of preventing Ras trafficking to the membrane and therefore inhibiting downstream signalling. This included generation of farnesyltransferase inhibitors, including lonafarnib and tipifarnib (Van Cutsem et al., 2004). However, the effectiveness of these was questionable, with most patients with KRAS-mutated diseases (such as pancreatic cancer and leukaemia patients) receiving no clinical benefit from these farnesyltransferase inhibitors (Borthakur et al., 2006; Burnett et al., 2012; Harousseau et al., 2009; Van Cutsem et al., 2004). Inefficacy was largely due to the redundancy mechanism of geranylgeranyltransferase, which sufficiently modifies KRAS in the absence of farnesyltransferase to permit its trafficking to the membrane (Basso et al., 2006; Whyte et al., 1997). Interest in this strategy has recently been revived with new personalised medicine approaches now capable of identifying patients harbouring HRAS- or NRAS-driven can- cers that are more likely to respond (Gilardi et al., 2020; Lee et al., 2020). In recent years however, more promising steps have been made regarding direct inhibition of oncogenic Ras. A key feature of this has been the discovery of novel potential binding pockets for small mole- cule inhibitors, to inhibit GEF activity or effector binding (Cruz-Migoni et al., 2019; Maurer et al., 2012; Ostrem et al., 2013). Fragment-based screening identified a previously undiscovered hydrophobic pocket located between the Switch I and II regions (termed S-IIP), which was successfully targeted by Ostrem et al. (2013) (Figure 4a). Binding of peptide fragments was specific to G12C-mutated KRAS since the compounds functioned through irreversible cysteine binding in this particular pocket (and not with other cysteines found in wild-type KRAS). Other key residues within this pocket include, but are not limited to, V7, V9, M72, F78, Q99 and I100. In vitro models of KRAS G12C-mutated lung cancer treated showed decreased survival upon treatment with these compounds, with inactive Ras (Ras-GDP) levels considerably greater than Ras-GTP. Further analysis showed that the conformational disruption caused by binding of these fragments reduced interactions with both SOS and effector molecules and pathways, including B-RAF, C-RAF and the PI3K pathway (Gentile et al., 2017; Ostrem et al., 2013). The binding of these fragments to Ras also reduce SOS-catalysed nucleotide exchange, a method of Ras inhibition which had been previously explored. Compounds acting in this way either inhibited conversion of Ras-GDP to Ras-GTP (Patgiri et al., 2011) or increased the amount of Ras-GTP to such a level that it inhibited ERK phosphor- ylation, since overactivation of Ras can be cytotoxic (the Ras “sweet- spot model”) (Li et al., 2018). Either way, these compounds were shown to inhibit the MAPK pathway but were largely tested against Ras wild type (in the context of inhibiting the effects of RTK muta- tions). Thus, the aforementioned compounds identified by Ostrem et al. were revolutionary in their specificity for targeting Ras-mutated disease. Discovery and characterisation of this SII-P pocket have led to the development of revolutionary KRAS G12C-selective covalent inhibitors, including ARS-853 and latterly AMG-510 (Figure 4b–d) (Canon et al., 2019; Patricelli et al., 2016). ARS-853 binds irreversibly to the inactive form of KRAS G12C, preventing exchange of GDP for GTP and therefore activation. This in turn inhibits downstream MAPK and PI3K–AKT pathway signalling, with KRAS–CRAF interactions sig- nificantly reduced. Moreover, in vitro evidence showed increased apo- ptosis and cell cycle arrest in some (but not all) models tested (Lito et al., 2016; Patricelli et al., 2016). This compound has subsequently been fully characterised in silico and these studies revealed a dynamic nature of the SII-P pocket, a feature which could be utilised in further study (Khrenova et al., 2020). Based on this work, alternative iterations of KRAS G12C inhibi- tors have been produced. This was required since the probability of ARS-853 locking KRAS in its inactive state in vivo was debatable, given a lack of understanding regarding the cycling efficiency of Ras between its inactive and active states. It had been deduced in vitro that the G12C mutation permits rapid cycling of KRAS between these states, hence permitting the binding of ARS-853 (Figure 4b). How- ever, the possibility of finding the correct therapeutic window to translate this compound to in vivo work proved complex (Janes et al., 2018; Lito et al., 2016; Patricelli et al., 2016). Thus, alternatives including ARS-1620 were developed to improve in vivo capability (Figure 4c). Modifications to the ARS-853 structure resulted in favourable pharmacokinetic (PK) properties, permitting a greater understanding of KRAS activation status and dependency in vivo (Janes et al., 2018). ARS-1620 is an orally bioavailable quinazoline- based compound with limited side effects witnessed in preclinical animal models (Janes et al., 2018; Li et al., 2018). Optimisation from ARS-853 by inclusion of a fluorophenol, hydrophobic binding group permitted stronger covalent (irreversible) binding within the SII-P pocket (more specifically, interaction with H95 in this pocket), thus 2854 HEALY ET AL. 14765381, 2022, 12, Downloaded from https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420, Wiley Online Library on [11/06/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
  • 12. FIGURE 4 Evolution of novel direct RAS-targeting agents. Chemical structure and protein structures of direct RAS-targeting agents in complex with GDP-bound (pink) KRAS. (a) Minor modification of initial compound hit 6H05, 6H05 compound 6 (purple) bound covalently to KRAS G12C, PDB accession no. 4LUC. (b) ARS-853 (purple) bound covalently to KRAS G12C (orange), PDB accession no. 5F2E. (c) ARS-1620 (purple) bound covalently to KRAS G12C (orange), PDB accession no. 5V9U. (d) AMG-510 (purple) bound to KRAS G12C (orange), PDB accession no. 6OIM. (e) MRTX849 (purple) bound covalently to KRAS G12C (orange), PDB accession no. 6UT0. Molecules shown in relation to the switch regions, largely binding in SII-P. All compounds bind covalently near to mutational hotspot HEALY ET AL. 2855 14765381, 2022, 12, Downloaded from https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420, Wiley Online Library on [11/06/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
  • 13. improving potency (Figure 4). The effects of this compound remained mutation specific, thereby eliminating the risk of binding to KRAS wild type in non-tumour cells and thus reduced toxicity potential. This compound was well tolerated in patient-derived xenograft mice, where a reduction in tumour burden through decreased Ras-mediated downstream signalling was evident. This was the first example of an in vivo trial using a compound targeting the SII-P pocket (Janes et al., 2018). From this, AMG-510 was developed, following further modifications to the ARS-1620 structure, including addition of more aromatic entities (Figure 4d). This enables AMG-510 to bind within a slightly different groove of SII-P, enhancing potency and selectivity (Canon et al., 2019). Promising preclinical in vitro and in vivo experiments indicated arrest of the MAPK pathway and induction of a pro-inflammatory tumour micro-environment. This compound has since progressed into clinical trials, the first KRAS G12C-selective inhibitor to do so. Early data indicate that out of 29 patients evaluable for response at the time of publication, 5 exhibited partial response, 18 had stable disease and 6 had progressive disease. As with the previous ARS-1620 animal studies, AMG-510 was generally well tolerated, with no dose limiting toxicities recorded (Govindan, 2019; Romero, 2020). However, larger cohorts and longer trials are imperative in determining the true impact of this compound. Such compounds could help combat the KRAS-mediated resistance to monoclonal antibodies seen in non-small cell lung cancer and colorec- tal cancer (Lièvre & Laurent-Puig, 2009; Park et al., 2019). However, success of this compound and indeed this targeting mechanism, is likely restricted to KRAS G12C-mutant cancer since some key residues for binding of this compound are unique to KRAS and not conserved between the different isoforms. Indeed, in vitro work completed by Ostrem et al. (2013) illustrated that transduction of lung cancer cell lines with KRAS G12V rescued the cancerous phenotype (resistance to cell death, increased proliferation), thereby illustrating how this mutation renders resistance to KRAS G12C inhibitors, as expected. Alternative KRAS G12C inhibitors are also in development and showing considerable promise, including MRTX849 (Fell et al., 2020; Hallin et al., 2020a). In a similar way to AMG-510 and the ARS com- pounds discussed above, MRTX849 covalently binds to the inactive form of KRAS, in SII-P (Figure 4e). This induces apoptosis through down-regulation of the MAPK pathway. Interestingly, the PI3K–AKT pathway remained relatively unaffected by MRTX849 (Hallin et al., 2020b). In vivo trials with MRTX849 exhibited favourable phar- macokinetic/pharmacodynamic properties (Fell et al., 2020) and both cell line- and patient-derived xenograft modelling of pancreatic and lung cancers also indicated up to a 30% reduction in tumour burden (Fell et al., 2020; Hallin et al., 2020b). Individual patient case studies from Phase I trials have also shown MRTX849 to be effective in reducing tumour burden in both lung cancer and colorectal cancer, although these data are largely incomplete (Hallin et al., 2020b). Taken together, it is clear that KRAS G12C inhibitors have promise as a means of abrogating Ras-mediated resistance, although there remain drawbacks which need assessing, most notably, the lack of efficacy against other Ras mutations, which are prevalent across Ras-mutated disease. Clearer understanding of the structure of Ras has permitted not only elucidation of the SII-P pocket but also alternative binding sites, including a pocket between the Switch I and II regions of Ras (termed pocket I). Key residues available for interaction with small molecule compounds include K5, L6, V7, V8, S39, D54, I55, L56, Y71, T74, G75 and E76 (Maurer et al., 2012; Quevedo et al., 2018). Antibody-frag- ment-directed site exploration can be used to explore and exploit pre- viously unconsidered drug interaction sites. In the case of Ras, this mechanism has been used to analyse potential compound binding sites within the previously identified pocket I that could be targeted using small molecule inhibitors to interrupt effector proteins (such as c-RAF and p110α) from binding to Ras via the Ras binding domain (Maurer et al., 2012; Quevedo et al., 2018). This would provide an alternative mechanism of abrogating the effects of oncogenic Ras activation to those discussed previously and early results showed effectiveness of antibody-derived compounds against a range of Ras mutations and isoforms (Quevedo et al., 2018). However, given high affinity binding of certain effectors to Ras (such as PI3K and B-RAF, with 3.2- and 0.04-μM affinity, respectively) (Erijman & Shifman, 2016), the high EC50s of the compounds identified in these in vitro assays mean that many further modifications would be required to convert these putative compounds into usable therapeu- tics. Nevertheless, such antibody-derived fragments have the poten- tial to be fused with small molecule protein–protein interaction inhibitors to improve efficacy (Cruz-Migoni et al., 2019). Whilst many Ras–effector interaction inhibitors have been trialled preclinically, none have been implemented in the clinic in the context of Ras- mutant cancer, owing to lack of efficacy or toxicity potential (Canon et al., 2019; Keeton et al., 2017). However, crystal structure determination showed that fusion of compounds developed through antibody-derived fragment screening with known small molecule protein–protein interaction inhibitors results in better binding within pocket I, thus inhibiting Ras–effector interactions with a lower EC50. Nevertheless, the therapeutic use of pocket I may be restricted since such a pocket has also been detected in wild-type Ras (Cruz-Migoni et al., 2019), thus increasing the risk for on-target toxicity. In recent times, inhibition of the Ras–effector interaction has been seen through competitive binding of rigosertib at the Ras bind- ing domain. This compound elicits effects against MAPK, PI3K and RAL pathway activation, in both wild-type and mutant Ras situations (Athuluri-Divakar et al., 2016). Inhibition of multiple Ras-mediated diseases have been seen in response to rigosertib, including pancre- atic cancer and leukaemia (Athuluri-Divakar et al., 2016; Baker et al., 2019). Rigosertib is moderately to well tolerated in clinical trials thus far, although is yet to be specifically tested in the context of Ras-mutant cancer (Bowles et al., 2014; Ma et al., 2012; Navada et al., 2020). As with the antibody-derived, small molecule protein– protein interaction inhibitors described above, toxicity may be as a result of rigosertib's ability to target both wild-type and mutant KRAS and thus, further studies are needed to fully assess the impact of such a drug on healthy cells. 2856 HEALY ET AL. 14765381, 2022, 12, Downloaded from https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420, Wiley Online Library on [11/06/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
  • 14. Antibody therapy is also currently being explored as means of direct Ras targeting. However, a major drawback has been the capa- bility of the antibody to cross the cell membrane, as with any protein- based therapy (Bolhassani et al., 2017). Therefore, the development of inRas37, a pan-Ras-targeting antibody, is a considerable step for- ward in targeting Ras. Although the cellular uptake remains low (approximately 4%), in vitro and in vivo work has shown promise in the potential of inRas37 to inhibit both the MAPK and PI3K–AKT path- ways, in a dose-dependent manner (Shin et al., 2020). Briefly, this drug binds to integrins αVβ3 and αVβ5 on the cancer cell surface which then undergo endocytosis. The antibody “escapes” the endosome as a result of pH-determined cleavage (the antibody is cleaved from the integrins better at pH 7, the cytoplasmic pH, compared to pH 6.5, the pH of early endosomes) (Podinovskaia & Spang, 2018; Putnam, 2012; Shin et al., 2020). The antibody then co-localises with Ras to block the effector binding site, in a manner similar to the protein–protein inter- action inhibitors described above. Mutations introduced into the gen- eral antibody structure render it specific for mutant Ras binding, with little activity against Ras wild type, thereby limiting on-target toxicity. inRas37 has greater effect on cells with a greater dependency on Ras signalling, which can include some tumour cells (Weinstein, 2002; Yi et al., 2020). Nevertheless, when tested on large tumour models (spheroids and cell-derived xenograft mice), efficacy decreased con- siderably (Shin et al., 2020). This therefore shows that treatment with this drug may be suitable for patients at earlier stages of Ras-mutated cancer, before tumour burden is too great and reduces drug efficacy. Thus, it could be used to treat Ras-initiated relapse as soon as it occurs. However, its utility in eliminating minor subclones prior to relapse is limited, given the low cellular uptake and the need for a high Ras dependency in the cell. 9 | RESISTANCE TO Ras-TARGETING AGENTS Of course, there is potential for resistance to any therapeutic and Ras-targeting drugs are no different. Some of these have been previ- ously discussed here, such as the use of geranylgeranyltransferase to overcome the effects of farnesyltransferase inhibitors (Whyte et al., 1997) or the up-regulation of alternative Ras-mediated path- ways, as seen when patients are treated with MEK inhibitors (Vitiello et al., 2019). Other mechanisms of resistance are also possible, such as the reactivation of ERK, which may be a potential resistance mechanism in the case of a Ras-targeting agent (Bruner et al., 2017; Ercan et al., 2012; Ochi et al., 2014). This has already been seen in the case of EGFR-inhibitor resistance, whereby negative regulators of ERK are down-regulated, so pro-apoptotic BIM is not fully up- regulated and so cannot fully induce apoptosis. Alternatively, the gene encoding ERK1, MAPK1, is amplified. These scenarios resulted in in vitro and in vivo resistance to the putative EGFR inhibitor WZ4002 (Ercan et al., 2012). ERK reactivation has also been found to contrib- ute to gefitinib resistance but, in this case, was found to be mediated by Src (Ochi et al., 2014). Given these kinases are either side of Ras, their co-operation could result in resistance to a Ras inhibitor. How- ever, Src-mediated ERK reactivation is avoidable through treatment with Src inhibitors (Ochi et al., 2014), which may present a means of overcoming this potential Ras-inhibitor resistance mechanism. Taken together, these studies imply that whilst resistance to Ras-targeting drugs is possible, there are already means of overcoming this resistance, just as a Ras-targeting drug would provide the means of overcoming Ras-mediated resistance. Whilst other elements of pathways contributing to resistance can be targeted relatively easily, acquisition of secondary mutations in Ras present a more pressing problem. For example, at present, the G12C- specific inhibitor is perhaps the most developed means of inhibiting Ras; however, a subsequent mutation in Ras would most likely render this inhibitor ineffective. This has already been evidenced in studies into KRAS G12C inhibitors, whereby rescue experiments with the G12V mutation restored the cancerous phenotype (Ostrem et al., 2013). Therefore, a pan-mutation-targeting drug, such as a derivative of inRas37, may be favoured. 10 | DISCUSSION Ras mutations in cancer and chemoresistance are important when considering patient prognosis. Whilst it seems inevitable that resis- tance will be an issue for a long time to come, better targeting of potential causes is imperative. Advances in Ras inhibition could help reduce the risk of resistance and relapse for a wide range of cancers, given its mutational frequency. At present, some success has been seen when multiple drugs are used to target different pathways impli- cated in Ras-mutated disease. However, other studies have shown lack of long-term efficacy when combining multiple therapies. There- fore, single agent, multi-pathway targeting agents, including direct Ras inhibitors, are becoming more heavily researched. It will be interesting to see the effects of these in vivo, since this type of therapy may reduce the potential for future development of drug resistance by up-regulation of an alternative pathway. Targeting a common factor at the centre of multiple pathways may target more cancer cell types and thus reduce the heterogeneity of the tumour, eliminating potential resistance causes before they become dominant. Previous failures of Ras-targeting agents, including farnesyl- transferase inhibitors, as well as perceived unfavourable protein dynamics, have resulted in Ras being largely considered undruggable. However, development of structural analysis techniques and a clearer understanding of the key residues in Ras has altered this thinking, with new compounds with novel, allele-specific Ras binding mechanisms showing great promise. Phase I trials of AMG-510, a first-in-class direct Ras-targeting agent, suggest a turning point has been reached in this field of study. Whilst there is much more to be done, these preliminary data indicate a solution for Ras-mediated resistance may be possible. Nevertheless, a greater understanding of resistance must be gained before relapse risk can be eliminated. Despite evidence supporting the cancer stem cell theory, limited standard-of-care detection sensitivities for initial diagnostic samples prevent HEALY ET AL. 2857 14765381, 2022, 12, Downloaded from https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420, Wiley Online Library on [11/06/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
  • 15. identification of the minor subclones present at diagnosis (McMahon et al., 2019). It can therefore be difficult to determine likely causes of relapse upon initial diagnosis and so constant monitoring for changes in expression of genes commonly implicated in drug resistance, such as Ras, may provide a useful tool for predicting disease trajectory. Nevertheless, this is only useful if the effects of the acquired/emer- gent mutations can be abrogated. In the cases discussed here, improved analytical tools would ideally be combined with Ras- targeting agents to prevent resistance taking hold. Ultimately, chemoresistance, either intrinsic or acquired, due to Ras mutations, whether primary or secondary, remains a considerable problem. The concepts presented here, amongst many other examples, illustrate the necessity for Ras-targeting drugs. There is a distinct clini- cal requirement for the improved targeting of Ras in cancer, with Ras implicated in both initial disease presentation and relapse. Although no universal, direct Ras inhibitor has yet been achieved, considerable pro- gress has been made in the last decade with the advent of allele-specific inhibitors. This brings promise to the field, with the potential for better treatment of Ras-initiated resistance a real prospect. 10.1 | Nomenclature of targets and ligands Key protein targets and ligands in this article are hyperlinked to corresponding entries in http://www.guidetopharmacology.org and are permanently archived in the Concise Guide to PHARMACOLOGY 2019/20 (Alexander, Kelly, Mathie, Peters, et al., 2019; Alexander, Fabbro, Kelly, Marrion, et al., 2019; Alexander, Fabbro, Kelly, Mathie, et al., 2019). CONFLICT OF INTEREST The authors declare no conflicts of interest. REFERENCES Adnane, L., Trail, P. A., Taylor, I., & Wilhelm, S. M. (2006). Sorafenib (BAY 43-9006, Nexavar®), a dual-action inhibitor that targets RAF/MEK/ERK pathway in tumor cells and tyrosine kinases VEGFR/PDGFR in tumor vasculature. In Methods in enzymology (pp. 597–612). Academic Press. Aikawa, T., Togashi, N., Iwanaga, K., Okada, H., Nishiya, Y., Inoue, S., Levis, M. J., & Isoyama, T. (2020). Quizartinib, a selective FLT3 inhibitor, maintains antileukemic activity in preclinical models of RAS-mediated midostaurin-resistant acute myeloid leukemia cells. Oncotarget, 11, 943–955. https://doi.org/10.18632/oncotarget. 27489 Alexander, S. P., Kelly, E., Mathie, A., Peters, J. A., Veale, E. L., Armstrong, J. F., Faccenda, E., Harding, S. D., Pawson, A. J., Sharman, J. L., Southan, C., Peter Buneman, O., Cidlowski, J. A., Christopoulos, A., Davenport, A. P., Fabbro, D., Spedding, M., Striessnig, J., Davies, J. A., & Collaborators, GTP. (2019). The concise guide to PHARMACOLOGY 2019/20: Introduction and Other Protein Targets. British Journal of Pharmacology, 176, S1–S20. https://doi.org/ 10.1111/bph.14747 Alexander, S. P. H., Fabbro, D., Kelly, E., Marrion, N. V., Mathie, A., Peters, J. A., Veale, E. L., Armstrong, J. F., Faccenda, E., Harding, S. D., Pawson, A. J., Sharman, J. L., Southan, C., Davies, J. A., & Collaborators, GTP. (2019). The concise guide to pharmacology 2019/20: Catalytic receptors. British Journal of Pharmacology, 176, S247–S296. https://doi.org/10.1111/bph.14751 Alexander, S. P. H., Fabbro, D., Kelly, E., Mathie, A., Peters, J. A., Veale, E. L., Armstrong, J. F., Faccenda, E., Harding, S. D., Pawson, A. J., Sharman, J. L., Southan, C., Davies, J. A., & GTP collabo- rators. (2019). The Concise Guide to PHARMACOLOGY 2019/20: Enzymes. British Journal of Pharmacology, 176, S297–S396. https:// doi.org/10.1111/bph.14752 Algazi, A. P., Esteve-Puig, R., Nosrati, A., Hinds, B., Hobbs- Muthukumar, A., Nandoskar, P., Ortiz-Urda, S., Chapman, P. B., & Daud, A. (2018). Dual MEK/AKT inhibition with trametinib and GSK2141795 does not yield clinical benefit in metastatic NRAS- mutant and wild-type melanoma. Pigment Cell & Melanoma Research, 31, 110–114. https://doi.org/10.1111/pcmr.12644 Al-Hajj, M., Wicha, M. S., Benito-Hernandez, A., Morrison, S. J., & Clarke, M. F. (2003). Prospective identification of tumorigenic breast cancer cells. Proceedings of the National Academy of Sciences, 100, 3983–3988. https://doi.org/10.1073/pnas.0530291100 Amado, R. G., Wolf, M., Peeters, M., Van Cutsem, E., Siena, S., Freeman, D. J., Juan, T., Sikorski, R., Suggs, S., Radinsky, R., Patterson, S. D., & Chang, D. D. (2008). Wild-type KRAS is required for panitumumab efficacy in patients with metastatic colorectal can- cer. Journal of Clinical Oncology, 26, 1626–1634. https://doi.org/10. 1200/JCO.2007.14.7116 Amirouchene-Angelozzi, N., Swanton, C., & Bardelli, A. (2017). Tumor evo- lution as a therapeutic target. Cancer Discovery, 7, 805–817. https:// doi.org/10.1158/2159-8290.CD-17-0343 André, F., Ciruelos, E., Rubovszky, G., Campone, M., Loibl, S., Rugo, H. S., et al. (2019). Alpelisib for PIK3CA-mutated, hormone receptor–positive advanced breast cancer. New England Journal of Medicine, 380, 1929–1940. https://doi.org/10.1056/NEJMoa 1813904 Athuluri-Divakar, S. K., Vasquez-Del Carpio, R., Dutta, K., Baker Stacey, J., Cosenza Stephen, C., Basu, I., Gupta, Y. K., Reddy, M. V. R., Ueno, L., Hart, J. R., Vogt, P. K., Mulholland, D., Guha, C., Aggarwal, A. K., & Reddy, E. P. (2016). A small molecule RAS-mimetic disrupts RAS asso- ciation with effector proteins to block signaling. Cell, 165, 643–655. https://doi.org/10.1016/j.cell.2016.03.045 Baker, S. J., Cosenza, S. C., Ramana Reddy, M. V., & Premkumar Reddy, E. (2019). Rigosertib ameliorates the effects of oncogenic KRAS signaling in a murine model of myeloproliferative neoplasia. Oncotarget, 10, 1932–1942. https://doi.org/10.18632/oncotarget.26735 Basak, P., Sadhukhan, P., Sarkar, P., & Sil, P. C. (2017). Perspectives of the Nrf-2 signaling pathway in cancer progression and therapy. Toxicology Reports, 4, 306–318. https://doi.org/10.1016/j.toxrep. 2017.06.002 Basso, A. D., Kirschmeier, P., & Bishop, W. R. (2006). Thematic review series: Lipid posttranslational modifications. Farnesyl transferase inhib- itors. Journal of Lipid Research, 47, 15–31. https://doi.org/10.1194/jlr. R500012-JLR200 Bhaduri, A., Di Lullo, E., Jung, D., Müller, S., Crouch, E. E., Espinosa, C. S., Ozawa, T., Alvarado, B., Spatazza, J., Cadwell, C. R., Wilkins, G., Velmeshev, D., Liu, S. J., Malatesta, M., Andrews, M. G., Mostajo- Radji, M. A., Huang, E. J., Nowakowski, T. J., Lim, D. A., … Kriegstein, A. R. (2020). Outer radial glia-like cancer stem cells contrib- ute to heterogeneity of glioblastoma. Cell Stem Cell, 26, 48–63.e46. https://doi.org/10.1016/j.stem.2019.11.015 Bhagwat, S. V., McMillen, W. T., Cai, S., Zhao, B., Whitesell, M., Shen, W., Kindler, L., Flack, R. S., Wu, W., Anderson, B., Zhai, Y., Yuan, X.-J., Pogue, M., Van Horn, Rao, X., McCann, D., Dropsey, A. J., Manro, J., Walgren, J., … Peng, S.-B. (2020). ERK inhibitor LY3214996 targets ERK pathway–driven cancers: A therapeutic approach toward preci- sion medicine. Molecular Cancer Therapeutics, 19, 325–336. https:// doi.org/10.1158/1535-7163.MCT-19-0183 2858 HEALY ET AL. 14765381, 2022, 12, Downloaded from https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420, Wiley Online Library on [11/06/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
  • 16. Bolhassani, A., Jafarzade, B. S., & Mardani, G. (2017). In vitro and in vivo delivery of therapeutic proteins using cell penetrating peptides. Pep- tides, 87, 50–63. https://doi.org/10.1016/j.peptides.2016.11.011 Bonnet, D., & Dick, J. E. (1997). Human acute myeloid leukemia is organized as a hierarchy that originates from a primitive hematopoi- etic cell. Nature Medicine, 3, 730–737. https://doi.org/10.1038/ nm0797-730 Borthakur, G., Kantarjian, H., Daley, G., Talpaz, M., O'Brien, S., Garcia- Manero, G., Giles, F., Faderl, S., Sugrue, M., & Cortes, J. (2006). Pilot study of lonafarnib, a farnesyl transferase inhibitor, in patients with chronic myeloid leukemia in the chronic or accelerated phase that is resistant or refractory to imatinib therapy. Cancer, 106, 346–352. https://doi.org/10.1002/cncr.21590 Borthakur, G., Popplewell, L., Boyiadzis, M., Foran, J., Platzbecker, U., Vey, N., Walter, R. B., Olin, R., Raza, A., Giagounidis, A., Al-Kali, A., Jabbour, E., Kadia, T., Garcia-Manero, G., Bauman, J. W., Wu, Y., Liu, Y., Schramek, D., Cox, D. S., … Kantarjian, H. (2016). Activity of the oral mitogen-activated protein kinase kinase inhibitor trametinib in RAS-mutant relapsed or refractory myeloid malignancies. Cancer, 122, 1871–1879. https://doi.org/10.1002/cncr.29986 Bowles, D. W., Diamond, J. R., Lam, E. T., Weekes, C. D., Astling, D. P., Anderson, R. T., Leong, S., Gore, L., Varella-Garcia, M., Vogler, B. W., Keysar, S. B., Freas, E., Aisner, D. L., Ren, C., Tan, A.-C., Wilhelm, F., Maniar, M., Eckhardt, S. G., Messersmith, W. A., & Jimeno, A. (2014). Phase I study of oral rigosertib (ON 01910.Na), a dual inhibitor of the PI3K and Plk1 pathways, in adult patients with advanced solid malig- nancies. Clinical Cancer Research, 20, 1656–1665. https://doi.org/10. 1158/1078-0432.CCR-13-2506 Bowyer, S., Lee, R., Fusi, A., & Lorigan, P. (2015). Dabrafenib and its use in the treatment of metastatic melanoma. Melanoma Manag, 2, 199–208. https://doi.org/10.2217/mmt.15.21 Boyer, L. A., Lee, T. I., Cole, M. F., Johnstone, S. E., Levine, S. S., Zucker, J. P., Guenther, M. G., Kumar, R. M., Murray, H. L., Jenner, R. G., Gifford, D. K., Melton, D. A., Jaenisch, R., & Young, R. A. (2005). Core transcriptional regulatory circuitry in human embryonic stem cells. Cell, 122, 947–956. https://doi.org/10.1016/j.cell.2005. 08.020 Bruner, J. K., Ma, H. S., Li, L., Qin, A. C. R., Rudek, M. A., Jones, R. J., Levis, M. J., Pratz, K. W., Pratilas, C. A., & Small, D. (2017). Adaptation to TKI treatment reactivates ERK signaling in tyrosine kinase–driven leukemias and other malignancies. Cancer Research, 77, 5554–5563. https://doi.org/10.1158/0008-5472.CAN-16-2593 Burnett, A. K., Russell, N. H., Culligan, D., Cavanagh, J., Kell, J., Wheatley, K., Virchis, A., Hills, R. K., Milligan, D., & AML Working Group of the UK National Cancer Research Institute. (2012). The addi- tion of the farnesyl transferase inhibitor, tipifarnib, to low dose cytarabine does not improve outcome for older patients with AML. British Journal of Haematology, 158, 519–522. https://doi.org/10. 1111/j.1365-2141.2012.09165.x Burrell, R. A., McGranahan, N., Bartek, J., & Swanton, C. (2013). The causes and consequences of genetic heterogeneity in cancer evolution. Nature, 501, 338–345. https://doi.org/10.1038/nature12625 Caiola, E., Salles, D., Frapolli, R., Lupi, M., Rotella, G., Ronchi, A., Garassino, M. C., Mattschas, N., Colavecchio, S., Broggini, M., Wiesmüller, L., & Marabese, M. (2015). Base excision repair-mediated resistance to cisplatin in KRAS(G12C) mutant NSCLC cells. Oncotarget, 6, 30072–30087. https://doi.org/10.18632/oncotarget.5019 Canon, J., Rex, K., Saiki, A. Y., Mohr, C., Cooke, K., Bagal, D., Gaida, K., Holt, T., Knutson, C. G., Koppada, N., Lanman, B. A., Werner, J., Rapaport, A. S., Miguel, T. S., Ortiz, R., Osgood, T., Sun, J.-R., Zhu, X., McCarter, J. D., … Lipford, J. R. (2019). The clinical KRAS(G12C) inhibi- tor AMG 510 drives anti-tumour immunity. Nature, 575, 217–223. https://doi.org/10.1038/s41586-019-1694-1 Cantley, L. C. (2002). The phosphoinositide 3-kinase pathway. Science, 296, 1655–1657. https://doi.org/10.1126/science.296.5573.1655 Castellano, E., & Downward, J. (2011). RAS interaction with PI3K: More than just another effector pathway. Genes & Cancer, 2, 261–274. https://doi.org/10.1177/1947601911408079 Cerami, E., Gao, J., Dogrusoz, U., Gross, B. E., Sumer, S. O., Aksoy, B. A., Jacobsen, A., Byrne, C. J., Heuer, M. L., Larsson, E., Antipin, Y., Reva, B., Goldberg, A. P., Sander, C., & Schultz, N. (2012). The cBio cancer genomics portal: An open platform for exploring multi- dimensional cancer genomics data. Cancer Discovery, 2, 401–404. https://doi.org/10.1158/2159-8290.CD-12-0095 Cha, P. H., Cho, Y. H., Lee, S. K., Lee, J., Jeong, W. J., Moon, B. S., Yun, J.- H., Yang, J. S., Choi, S., Yoon, J., Kim, H.-Y., Kim, M.-Y., Kaduwal, S., Lee, W., Min, D. S., Kim, H., Han, G., & Choi, K.-Y. (2016). Small- molecule binding of the axin RGS domain promotes β-catenin and Ras degradation. Nature Chemical Biology, 12, 593–600. https://doi.org/ 10.1038/nchembio.2103 Chan, L. H., Zhou, L., Ng, K. Y., Wong, T. L., Lee, T. K., Sharma, R., Loong, J. H., Ching, Y. P., Yuan, Y.-F., Xie, D., Lo, C. M., Man, K., Artegiani, B., Clevers, H., Yan, H. H., Leung, S. Y., Richard, S., Guan, X.- Y., Huen, M. S. Y., & Ma, S. (2018). PRMT6 regulates RAS/RAF binding and MEK/ERK-mediated cancer stemness activities in hepatocellular carcinoma through CRAF methylation. Cell Reports, 25, 690–701e698. https://doi.org/10.1016/j.celrep.2018.09.053 Chang, F., Lee, J. T., Navolanic, P. M., Steelman, L. S., Shelton, J. G., Blalock, W. L., Franklin, R. A., & McCubrey, J. A. (2003). Involvement of PI3K/Akt pathway in cell cycle progression, apoptosis, and neoplas- tic transformation: A target for cancer chemotherapy. Leukemia, 17, 590–603. https://doi.org/10.1038/sj.leu.2402824 Cho, Y.-H., Ro, E. J., Yoon, J.-S., Kwak, D.-K., Cho, J., Kang, D. W., Lee, H.- Y., & Choi, K.-Y. (2020). Small molecule-induced simultaneous destabilization of β-catenin and RAS is an effective molecular strategy to suppress stemness of colorectal cancer cells. Cell Communication and Signaling: CCS, 18, 38. https://doi.org/10.1186/s12964-020- 0519-z Ciardiello, F., & Tortora, G. (2008). EGFR antagonists in cancer treatment. New England Journal of Medicine, 358, 1160–1174. https://doi.org/10. 1056/NEJMra0707704 Corces-Zimmerman, M. R., Hong, W. J., Weissman, I. L., Medeiros, B. C., & Majeti, R. (2014). Preleukemic mutations in human acute myeloid leukemia affect epigenetic regulators and per- sist in remission. Proceedings of the National Academy of Sciences of the United States of America, 111, 2548–2553. https://doi.org/10. 1073/pnas.1324297111 Cox, A. D., & Der, C. J. (2002). Farnesyltransferase inhibitors: Promises and realities. Current Opinion in Pharmacology, 2, 388–393. https://doi. org/10.1016/S1471-4892(02)00181-9 Cox, A. D., Fesik, S. W., Kimmelman, A. C., Luo, J., & Der, C. J. (2014). Drugging the undruggable RAS: Mission possible? Nature Reviews Drug Discovery, 13, 828–851. https://doi.org/10.1038/nrd4389 Cruz-Migoni, A., Canning, P., Quevedo, C. E., Bataille, C. J. R., Bery, N., Miller, A., Russell, A. J., Phillips, S. E. V., Carr, S. B., & Rabbitts, T. H. (2019). Structure-based development of new RAS-effector inhibitors from a combination of active and inactive RAS-binding compounds. Proceedings of the National Academy of Sciences, 116, 2545–2550. https://doi.org/10.1073/pnas.1811360116 D'Abaco, G. M., Whitehead, R. H., & Burgess, A. W. (1996). Synergy between Apc min and an activated ras mutation is sufficient to induce colon carcinomas. Molecular and Cellular Biology, 16, 884–891. https:// doi.org/10.1128/MCB.16.3.884 De Roock, Claes, B., Bernasconi, D., De Schutter, Biesmans, B., Fountzilas, G., Kalogeras, K. T., Kotoula, V., Papamichael, D., Laurent- Puig, P., Penault-Llorca, F., Rougier, P., Vincenzi, B., Santini, D., Tonini, G., Cappuzzo, F., Frattini, M., Molinari, F., Saletti, P., … Tejpar, S. (2010). Effects of KRAS, BRAF, NRAS, and PIK3CA muta- tions on the efficacy of cetuximab plus chemotherapy in chemotherapy-refractory metastatic colorectal cancer: A retrospective HEALY ET AL. 2859 14765381, 2022, 12, Downloaded from https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420, Wiley Online Library on [11/06/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
  • 17. consortium analysis. The Lancet Oncology, 11, 753–762. https://doi. org/10.1016/S1470-2045(10)70130-3 de Vries, G., Rosas-Plaza, X., van Vugt, M. A. T. M., Gietema, J. A., & de Jong, S. (2020). Testicular cancer: Determinants of cisplatin sensitivity and novel therapeutic opportunities. Cancer Treatment Reviews, 88, 102054. https://doi.org/10.1016/j.ctrv.2020.102054 Dean, M., Fojo, T., & Bates, S. (2005). Tumour stem cells and drug resis- tance. Nature Reviews Cancer, 5, 275–284. https://doi.org/10.1038/ nrc1590 DeNicola, G. M., Karreth, F. A., Humpton, T. J., Gopinathan, A., Wei, C., Frese, K., Mangal, D., Yu, K. H., Yeo, C. J., Calhoun, E. S., Scrimieri, F., Winter, J. M., Hruban, R. H., Iacobuzio-Donahue, C., Kern, S. E., Blair, I. A., & Tuveson, D. A. (2011). Oncogene-induced Nrf2 transcrip- tion promotes ROS detoxification and tumorigenesis. Nature, 475, 106–109. https://doi.org/10.1038/nature10189 Desai, P., Mencia-Trinchant, N., Savenkov, O., Simon, M. S., Cheang, G., Lee, S., Samuel, M., Ritchie, E. K., Guzman, M. L., Ballman, K. V., Roboz, G. J., & Hassane, D. C. (2018). Somatic mutations precede acute myeloid leukemia years before diagnosis. Nature Medicine, 24, 1015–1023. https://doi.org/10.1038/s41591-018-0081-z Diaz, L. A. Jr., Williams, R. T., Wu, J., Kinde, I., Hecht, J. R., Berlin, J., Allen, B., Bozic, I., Reiter, J. G., Nowak, M. A., Kinzler, K. W., Oliner, K. S., & Vogelstein, B. (2012). The molecular evolution of acquired resistance to targeted EGFR blockade in colorectal cancers. Nature, 486, 537–540. https://doi.org/10.1038/nature11219 Du, R., Shen, W., Liu, Y., Gao, W., Zhou, W., Li, J., Zhao, S., Chen, C., Chen, Y., Liu, Y., Sun, P., Xiang, R., Shi, Y., & Luo, Y. (2019). TGIF2 pro- motes the progression of lung adenocarcinoma by bridging EGFR/RAS/ERK signaling to cancer cell stemness. Signal Transduction and Targeted Therapy, 4, 60. https://doi.org/10.1038/s41392-019- 0098-x Duffy, M. J., & Crown, J. (2021). Drugging “undruggable” genes for cancer treatment: Are we making progress? International Journal of Cancer, 148, 8–17. https://doi.org/10.1002/ijc.33197 Dunn, C., & Croom, K. F. (2006). Everolimus. Drugs, 66, 547–570. https:// doi.org/10.2165/00003495-200666040-00009 Duronio, V. (2008). The life of a cell: Apoptosis regulation by the PI3K/PKB pathway. Biochemical Journal, 415, 333–344. https://doi. org/10.1042/BJ20081056 Eberlein, C. A., Stetson, D., Markovets, A. A., Al-Kadhimi, K. J., Lai, Z., Fisher, P. R., Meador, C. B., Spitzler, P., Ichihara, E., Ross, S. J., Ahdesmaki, M. J., Ahmed, A., Ratcliffe, L. E., Christey O'Brien, E. L., Barnes, C. H., Brown, H., Smith, P. D., Dry, J. R., Beran, G., … Cross, D. A. E. (2015). Acquired resistance to the mutant-selective EGFR inhibitor AZD9291 is associated with increased dependence on RAS signaling in preclinical models. Cancer Research, 75, 2489–2500. https://doi.org/10.1158/0008-5472.CAN-14-3167 Engelman, J. A., Luo, J., & Cantley, L. C. (2006). The evolution of pho- sphatidylinositol 3-kinases as regulators of growth and metabolism. Nature Reviews Genetics, 7, 606–619. https://doi.org/10.1038/ nrg1879 Ercan, D., Xu, C., Yanagita, M., Monast, C. S., Pratilas, C. A., Montero, J., Butaney, M., Shimamura, T., Sholl, L., Ivanova, E. V., Tadi, M., Rogers, A., Repellin, C., Capelletti, M., Maertens, O., Goetz, E. M., Letai, A., Garraway, L. A., Lazzara, M. J., … Jänne, P. A. (2012). Reac- tivation of ERK signaling causes resistance to EGFR kinase inhibitors. Cancer Discovery, 2, 934–947. https://doi.org/10.1158/2159-8290. CD-12-0103 Erijman, A., & Shifman, J. M. (2016). RAS/effector interactions from structural and biophysical perspective. Mini Reviews in Medicinal Chemistry, 16, 370–375. https://doi.org/10.2174/ 1389557515666151001141838 Evelyn, C. R., Biesiada, J., Duan, X., Tang, H., Shang, X., Papoian, R., Seibel, W. L., Nelson, S., Meller, J., & Zheng, Y. (2015). Combined ratio- nal design and a high throughput screening platform for identifying chemical inhibitors of a Ras-activating enzyme. Journal of Biological Chemistry, 290, 12879–12898. https://doi.org/10.1074/jbc.M114. 634493 Evelyn, C. R., Duan, X., Biesiada, J., Seibel William, L., Meller, J., & Zheng, Y. (2014). Rational design of small molecule inhibitors targeting the Ras GEF, SOS1. Chemistry & Biology, 21, 1618–1628. https://doi. org/10.1016/j.chembiol.2014.09.018 Feldman, D. R., Iyer, G., Van Alstine, L., Patil, S., Al-Ahmadie, H., Reuter, V. E., Bosl, G. J., Chaganti, R. S., & Solit, D. B. (2014). Presence of somatic mutations within PIK3CA, AKT, RAS, and FGFR3 but not BRAF in cisplatin-resistant germ cell tumors. Clinical Cancer Research, 20, 3712–3720. https://doi.org/10.1158/1078-0432.CCR- 13-2868 Fell, J. B., Fischer, J. P., Baer, B. R., Blake, J. F., Bouhana, K., Briere, D. M., Brown, K. D., Burgess, L. E., Burns, A. C., Burkard, M. R., Chiang, H., Chicarelli, M. J., Cook, A. W., Gaudino, J. J., Hallin, J., Hanson, L., Hartley, D. P., Hicken, E. J., Hingorani, G. P., … Marx, M. A. (2020). Identification of the clinical development candidate MRTX849, a covalent KRASG12C inhibitor for the treatment of cancer. Journal of Medicinal Chemistry, 63, 6679–6693. https://doi.org/10.1021/acs. jmedchem.9b02052 Ferino, A., Rapozzi, V., & Xodo, L. E. (2020). The ROS-KRAS-Nrf2 axis in the control of the redox homeostasis and the intersection with survival-apoptosis pathways: Implications for photodynamic therapy. Journal of Photochemistry and Photobiology B: Biology, 202, 111672. https://doi.org/10.1016/j.jphotobiol.2019.111672 Freeman, W. H. (2000). Section 20.4 receptor tyrosine kinases and RAS. In H. B. A. Lodish, S. L. Zipursky, et al. (Eds.), Molecular cell biology (4th ed.). New York: W.H. Freeman. Furet, P., Guagnano, V., Fairhurst, R. A., Imbach-Weese, P., Bruce, I., Knapp, M., Fritsch, C., Blasco, F., Blanz, J., Aichholz, R., Hamon, J., Fabbro, D., & Caravatti, G. (2013). Discovery of NVP-BYL719 a potent and selective phosphatidylinositol-3 kinase alpha inhibitor selected for clinical evaluation. Bioorganic & Medicinal Chemistry Letters, 23, 3741–3748. https://doi.org/10.1016/j.bmcl.2013.05.007 Galanis, A., Ma, H., Rajkhowa, T., Ramachandran, A., Small, D., Cortes, J., & Levis, M. (2014). Crenolanib is a potent inhibitor of FLT3 with activity against resistance-conferring point mutants. Blood, 123, 94–100. https://doi.org/10.1182/blood-2013-10-529313 Gao, J., Aksoy, B. A., Dogrusoz, U., Dresdner, G., Gross, B., Sumer, S. O., Sun, Y., Jacobsen, A., Sinha, R., Larsson, E., Cerami, E., Sander, C., & Schultz, N. (2013). Integrative analysis of complex cancer genomics and clinical profiles using the cBioPortal. Science Signaling, 6, pl1. https://doi.org/10.1126/scisignal.2004088 Garassino, M. C., Marabese, M., Rusconi, P., Rulli, E., Martelli, O., Farina, G., Scanni, A., & Broggini, M. (2011). Different types of K-Ras mutations could affect drug sensitivity and tumour behaviour in non- small-cell lung cancer. Annals of Oncology, 22, 235–237. https://doi. org/10.1093/annonc/mdq680 Garnock-Jones, K. P. (2015). Cobimetinib: First global approval. Drugs, 75, 1823–1830. https://doi.org/10.1007/s40265-015-0477-8 Gentile, D. R., Rathinaswamy, M. K., Jenkins, M. L., Moss, S. M., Siempelkamp, B. D., Renslo, A. R., Burke, J. E., & Shokat, K. M. (2017). Ras binder induces a modified switch-II pocket in GTP and GDP states. Cell Chemical Biology, 24, 1455–1466e1414. https://doi.org/10.1016/ j.chembiol.2017.08.025 Gerlach, D., Gmachl, M., Ramharter, J., Teh, J., Fu, S.-C., Trapani, F., Kessler, D., Rumpel, K., Botesteanu, D.-A., Ettmayer, P., Arnhof, H., Gerstberger, T., Kofink, C., Wunberg, T., Vellano, C. P., Heffernan, T. P., Marszalek, J. R., Pearson, M., McConnell, D. B., … Hofmann, M. H. (2020). Abstract 1091: BI-3406 and BI 1701963: Potent and selective SOS1::KRAS inhibitors induce regressions in combination with MEK inhibitors or irinotecan. Cancer Research, 80, 1091. 2860 HEALY ET AL. 14765381, 2022, 12, Downloaded from https://bpspubs.onlinelibrary.wiley.com/doi/10.1111/bph.15420, Wiley Online Library on [11/06/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License